comparison between satellite-based and cosmic ray …

160
COMPARISON BETWEEN SATELLITE-BASED AND COSMIC RAY PROBE SOIL MOISTURE ESTIMATES: A CASE STUDY IN THE CATHEDRAL PEAK CATCHMENT THIGESH VATHER Submitted in partial fulfilment of the requirements for the degree of MSc in Hydrology School of Agriculture, Earth and Environmental Sciences University of KwaZulu-Natal Pietermaritzburg Supervisor: Ms K T Chetty Co-supervisors: Dr M G Mengistu Prof C S Everson November 2015

Upload: others

Post on 26-Dec-2021

0 views

Category:

Documents


0 download

TRANSCRIPT

COMPARISON BETWEEN SATELLITE-BASED AND COSMIC RAY PROBE SOIL MOISTURE ESTIMATES: A

CASE STUDY IN THE CATHEDRAL PEAK CATCHMENT

THIGESH VATHER

Submitted in partial fulfilment of the requirements for the degree of MSc in Hydrology

School of Agriculture, Earth and Environmental Sciences

University of KwaZulu-Natal

Pietermaritzburg

Supervisor: Ms K T Chetty

Co-supervisors: Dr M G Mengistu Prof C S Everson

November 2015

i

ABSTRACT

Soil moisture is an important hydrological parameter, which is essential for a variety of

applications, extending to numerous disciplines. Currently, there are three methods of

estimating soil moisture. These include: (a) ground-based (in-situ) measurements, which are

carried out using field instruments; (b) remote sensing based methods, which use specialized

sensors on satellites and aircrafts and (c) land surface models, which use meteorological data

as inputs, at a predefined spatial resolution (Albergel et al. (2012); Mecklenburg et al., 2013).

In recent years the cosmic ray probe (CRP), which is an in-situ technique, has been

implemented in several countries across the globe. The CRP provides area-averaged soil

moisture at an intermediate scale and thus bridges the gap between in-situ point

measurements and satellite-based soil moisture estimates (Zreda et al., 2012). The aim of this

study was to first evaluate the current techniques for soil moisture estimation, in order to

identify the research gaps and limitations. The key objectives of this study were to test the

suitability of the CRP to provide spatial estimates of soil moisture and use these estimates to

validate satellite-based (remote sensing and modelled) soil moisture estimates in the

Cathedral Peak Catchment VI. The CRP was set-up and calibrated in Cathedral Peak

Catchment VI. An in-situ soil moisture network was created in Catchment VI, which was

used to validate the calibrated CRP soil moisture estimates. Once calibrated, the CRP was

found to provide spatial estimates of soil moisture, which correlated well with the in-situ soil

moisture network dataset and yielded a R2 value of 0.8445. The calibrated CRP was used to

validate satellite-based soil moisture products. The remote sensing products used were the

Level Three AMSR2 and SMOS products. The AMSR2 and SMOS products generally

under-estimated soil moisture throughout, but followed the general trend of the CRP, with

AMSR obtaining a R2 of 0.505 and SMOS obtaining a R2 of 0.4853, when compared against

the CRP estimates. The CRP was used to validate modelled soil moisture products, which

consisted of the SAHG product and the back-calculation of soil moisture, using equations by

Su et al. (2003) and Scott et al. (2003), and products derived from the SEBS Model. The

SAHG Model performed well, as it provided estimates that correlated well with the CRP

dataset and yielded a R2 value of 0.624 compared to the CRP estimates. The SEBS back-

calculation technique performed very poorly, as it over-estimated in the wet periods and

under-estimated in the dry periods.

ii

DECLARATION

I, Thigesh Vather, declare that:

(i) the research reported in this dissertation, except where otherwise indicated, is my

original work.

(ii) this dissertation has not been submitted for any degree or examination at any other

university.

(iii) this dissertation does not contain other persons’ data, pictures, graphs or other

information, unless specifically acknowledged as being sourced from other persons.

(iv) this dissertation does not contain other persons’ writing, unless specifically

acknowledged as being sourced from other researchers. Where other written sources

have been quoted, then:

(a) their words have been re-written, but the general information attributed to them

has been referenced;

(b) where their exact words have been used, their writing has been placed inside

quotation marks, and referenced.

(v) Where I have reproduced a publication of which I am an author, co-author or editor, I

have indicated, in detail, which part of the publication was actually written by myself

alone and have fully referenced such publications.

(vi) This dissertation does not contain text, graphics or tables copied and pasted from the

Internet, unless specifically acknowledged, and the source being detailed in the

Dissertation and in the References sections.

Signed: ………………..

Thigesh Vather

Supervisor: …………………………..

Miss Kershani Tinisha Chetty

Co-supervisor: …………………………

Dr Michael Mengistu

iii

PREFACE

The work undertaken and described in this dissertation was carried out in the Centre for

Water Resources Research (CWRR), in the School of Agriculture, Earth and Environmental

Science. This school is within the University of KwaZulu-Natal, Pietermaritzburg. The

following dissertation was supervised by Ms KT Chetty, Dr MG Mengistu and Prof CS

Everson.

This dissertation is original, unpublished, independent work by the author Thigesh Vather.

The work of other authors, when used, has been given the appropriate credit.

iv

ACKNOWLEDGEMENTS

The following Masters Research titled “Comparison Between Satellite-Based and Cosmic

Ray Probe Soil Moisture Estimates: A Case Study in the Cathedral Peak Catchment’’ has

been funded by the Centre for Water Resources Research (CWRR) and the National Research

Foundation (NRF). I am grateful for the aforementioned institutions for the funding they have

contributed towards this research. I would also like to acknowledge the following people and

institutions for their fundamental role in the completion of this research project:

First and foremost, I would like to thank my parents and sisters for their continuous love and

support. I appreciate all that you have done for me, not only during the past two years, but

throughout my life Thank you to my extended family and friends for all the love and support

you have given me. Thank you to Vadinie Moodley, for all the love and support, as well as

keeping me focused throughout my Masters.

Miss KT Chetty (supervisor). Thank you for your guidance and support throughout

the duration of the research project. Your mentorship has been of much value to the

completion of the dissertation.

Dr MG Mengistu (co-supervisor). Thank you for all the support and mentorship. Your

contribution towards the completion of this research study has been invaluable. Thank

you for all the field trips to Cathedral Peak and assisting me will all the aspects of my

project, such as the setting up of the cosmic ray probe, setting up of the soil moisture

network, calibrating the cosmic ray probe, teaching me how to set-up and run the

SEBS model and for assisting with any problems that arose.

Prof CS Everson (co-supervisor). Thank you for all the support and mentorship

throughout the project. Thank you for making this research project possible by

obtaining the cosmic ray probe. Thank you for all the well organized and productive

field visits. Thank you for assisting me with the various fieldwork activities that were

carried out. Thank you for organizing the various instrumentation and support

throughout the project.

v

Mrs S Rees. Thank you for assisting me with the editing of the dissertation. I really

appreciate the time and effort that you have put into improving the dissertation.

Mr S Mfeka and Mr C Pretorius. Thank you for all the assistance in the field and the

laboratory. Thank you for always being willing to help and always approachable.

The winter school. Thank you for assisting me with the first calibration, which was

the most labour intensive and time consuming calibration. I appreciate all the work

and effort that was made.

Prof TE Franz. Thank you for assisting me with the third calibration, as well as

teaching me the cosmic ray probe calibration procedure.

Dr S Sinclair. Thank you for the assistance in providing the PyTOPKAPI soil

moisture product.

Mr S Gokool. Thank you for your assistance and advice throughout the duration of

the project. Your suggestions towards improving the project were greatly beneficial

and appreciated.

Mr B Scott-Shaw. Thank you for assisting with the fire breaks.

Mrs S van Rensburg. Thank you for organizing the accommodation and other aspects

of the field trip. The work and effort that you have contributed is greatly appreciated.

Thank you to the Staff and Students of the Hydrology Department for all the help and

continuous support.

vi

TABLE OF CONTENTS Page

ABSTRACT ................................................................................................................................ i

PREFACE ................................................................................................................................. iii

ACKNOWLEDGEMENTS ...................................................................................................... iv

LIST OF TABLES .................................................................................................................... ix

LIST OF FIGURES ....................................................................................................................x

LIST OF SYMBOLS .............................................................................................................. xvi

LIST OF ABBREVIATIONS ................................................................................................. xix

1. INTRODUCTION ..............................................................................................................1

2. IN-SITU METHODS OF SOIL MOISTURE MEASUREMENT .....................................4

2.1 Conventional Methods of Soil Moisture Estimation ..................................................4

2.2 Cosmic Ray Probe ......................................................................................................6

Production of cosmic ray neutrons ...................................................................7

Moderation of neutrons ....................................................................................8

Cosmic ray probe measurements ......................................................................8

Measurement footprint and depth of the cosmic ray probe ..............................9

Advantages of the cosmic ray probe method .................................................11

3. REMOTE SENSING OF SOIL MOISTURE ...................................................................14

3.1 Overview of Remote Sensing of Soil Moisture ........................................................14

3.2 AMSR2 .....................................................................................................................15

3.3 SMOS ........................................................................................................................17

vii

3.4 Downscaling Techniques ..........................................................................................21

4. METHODS OF MODELLING SOIL MOISTURE .........................................................25

4.1 Land Surface Model (PyTOPKAPI) .........................................................................25

4.2 Surface Energy Balance System (SEBS) ..................................................................29

5. SYNTHESIS OF LITERATURE .....................................................................................34

6. METHODOLOGY ...........................................................................................................36

6.1 Study Site ..................................................................................................................37

6.2 Cosmic Ray Probe ....................................................................................................39

Set up of the cosmic ray probe .......................................................................39

Field sampling ................................................................................................42

Gravimetric water content determination .......................................................42

Creating a soil moisture network ....................................................................43

Catchment burning .........................................................................................44

Bulk density determination ............................................................................46

6.3 Calibration ................................................................................................................48

6.4 Creating an In-situ Soil Moisture Dataset .................................................................60

6.5 Acquisition and Processing of the AMSR2 Soil Moisture Product ..........................64

6.6 Acquisition and Processing of the SMOS Soil Moisture Product ............................67

6.7 Analysis of AMSR2 and SMOS Remote Sensing Data ...........................................71

6.8 PyTOPKAPI Soil Moisture Product (SAHG) ..........................................................74

6.9 Surface Energy Balance System (SEBS) ..................................................................80

viii

7. RESULTS AND DISCUSSION .......................................................................................94

7.1 Validating the CRP ...................................................................................................95

7.2 AMSR2 Soil Moisture Product Validation ...............................................................99

7.3 SMOS Soil Moisture Product Validation ...............................................................106

7.4 Comparing Remote Sensing Soil Moisture Products .............................................111

7.5 SAHG Soil Moisture Product Validation ...............................................................113

7.6 SEBS Soil Moisture Validation ..............................................................................119

7.7 Evaluating the Satellite-Based Soil Moisture Products ..........................................128

8. CONCLUSIONS AND RECOMMENDATIONS .........................................................130

8.1 Conclusions .............................................................................................................130

8.2 Recommendations ...................................................................................................132

9. REFERENCES ...............................................................................................................134

ix

LIST OF TABLES Page

Table 2.1 Advantages of the CRP ......................................................................................... 11

Table 3.1 Evaluation of downscaling techniques .................................................................. 22

Table 6.1 Geographical coordinates of sample points .......................................................... 41

Table 6.2 Variables required in obtaining the bulk density .................................................. 48

Table 6.3 Information regarding the calibration sampling.................................................... 49

Table 6.4 The gravimetric soil moisture, bulk density and Neutron count ........................... 57

Table 6.5 Date and calculated No value for each calibration ................................................ 58

Table 6.6 The bulk density of Catchment VI, as determined by Everson et al. (1998) ........ 79

Table 6.7 Mean bulk density according to depth .................................................................. 79

Table 6.8 Characteristics of the various landsat 8 bands (USGS, 2015). ............................. 81

Table 6.9 Calculated ESUN values ....................................................................................... 84

Table 6.10 Zonal statistics of relative evaporation enclosed by catchment VI ...................... 92

Table 7.1 T-test of In-situ against CRP estimates ................................................................. 98

Table 7.2 T-test of CRP against AMSR2 estimates ............................................................ 103

Table 7.3 T-test of CRP against SMOS estimates .............................................................. 110

Table 7.4 T-test of CRP against SAHG estimates .............................................................. 118

Table 7.5 Relative evaporation and evaporative fraction calculated using the SEBS

model for Catchment VI ..................................................................................... 120

Table 7.6 Evaporative fraction estimates from the SEBS Model and eddy covariance

technique ............................................................................................................. 126

Table 7.7 Spatial resolution of soil moisture products ........................................................ 128

x

LIST OF FIGURES

Figure 2.1 CRP system in Cathedral Peak Catchment VI ........................................................ 7

Figure 2.2 Difference in neutron concentration according to soil moisture content

(Franz et al., 2012b). ............................................................................................... 9

Figure 2.3 Measurement area and depth of CRP (Ochsner et al., 2013) ................................ 10

Figure 3.1 AMSR2 sensor on-board the GCOM-W1 satellite ............................................... 16

Figure 3.2 SMOS satellite ...................................................................................................... 18

Figure 4.1 A schematic of the water transfer in a typical PyTOPKAPI model cell

(Sinclair and Pegram, 2012) .................................................................................. 26

Figure 4.2 Data-flow diagram showing sources of the dynamic and static data to

produce the main information streams (Sinclair and Pegram, 2010). ................... 27

Figure 6.1 Location of Cathedral Peak Catchment VI, within the Tugela catchment,

within KwaZulu-Natal .......................................................................................... 37

Figure 6.2 Topographic map of Catchment VI ...................................................................... 38

Figure 6.3 CRP in Cathedral Peak Catchment VI .................................................................. 39

Figure 6.4 Diagram of sampling points .................................................................................. 40

Figure 6.5 Position of calibration sample points within Cathedral Peak Catchment VI ........ 41

Figure 6.6 Field samples contained in plastic bags ................................................................ 42

Figure 6.7 Weighing the soil samples and placing them in the oven ..................................... 43

Figure 6.8 (a) Setting up the TDR pit, (b) Wireless TDR, (c) Echo probe and (d) Data

Hobo Onset logger for the Echo probe ................................................................. 44

Figure 6.9 Data retrieved from burned echo probe data loggers ............................................ 46

Figure 6.10 Protection of equipment by fire breaks in Catchment VI ..................................... 46

xi

Figure 6.11 Soil moisture map of the first calibration (9th of July 2014) ................................ 50

Figure 6.12 Soil moisture map of the second calibration (28th of August 2014) ..................... 50

Figure 6.13 Soil moisture map of the third calibration (2nd of December 2014) ..................... 51

Figure 6.14 Soil moisture map of the fourth calibration (22nd of January 2015) ..................... 51

Figure 6.15 Volumetric water content against depth for each replicate at one sample

point ...................................................................................................................... 52

Figure 6.16 Volumetric water content against depth for all 24 sample points ......................... 53

Figure 6.17 Hourly temperature (The same data gaps present in the relative humidity

dataset) .................................................................................................................. 55

Figure 6.18 Complete daily air temperature data for Cathedral Peak Catchment VI ............... 56

Figure 6.19 CRP soil moisture estimates prior to calibration, with calibration points ............ 58

Figure 6.20 Hourly soil moisture estimates of the calibrated CRP .......................................... 59

Figure 6.21 Neutron count against volumetric water content .................................................. 59

Figure 6.22 Daily calibrated CRP soil moisture measurements in Catchment VI ................... 60

Figure 6.23 Daily TDR pit soil moisture measurements in Catchment VI .............................. 61

Figure 6.24 The CRP effective measurement depth ................................................................. 62

Figure 6.25 The average TDR pit soil moisture measurements in Catchment VI ................... 62

Figure 6.26 Daily wireless TDR data in Catchment VI ........................................................... 63

Figure 6.27 Daily echo probe data in Catchment VI ................................................................ 63

Figure 6.28 Representative soil moisture measurements ......................................................... 64

Figure 6.29 Comparison of spatial resolution of AMSR2 10 km and 25 km Level Three

soil moisture products ........................................................................................... 65

xii

Figure 6.30 The Catchment VI shapefile overlaid over the AMSR2 soil moisture

product to obtain the volumetric water content .................................................... 66

Figure 6.31 Navigation through the THREDDS service .......................................................... 68

Figure 6.32 Layout of the Godiva2 online visualization tool ................................................... 69

Figure 6.33 The position of the catchment in relation to a single pixel of the SMOS soil

moisture product ................................................................................................... 70

Figure 6.34 Godiva2 visualization tool showing the pixel value for the Catchment VI .......... 71

Figure 6.35 AMSR2 data availability for the observation period ............................................ 72

Figure 6.36 SMOS data availability for the observation period ............................................... 72

Figure 6.37 Time series analysis of days which have both ascending and descending

AMSR2 values ...................................................................................................... 73

Figure 6.38 Time series analysis of days which have both ascending and descending

SMOS values......................................................................................................... 73

Figure 6.39 Navigation through the SAHG website to download SSI data ............................. 75

Figure 6.40 One day of SSI data (eight images on a three hour interval) ................................ 76

Figure 6.41 Daily SAHG soil moisture product ....................................................................... 77

Figure 6.42 Overlay of the Cathedral Peak Catchment VI onto the SAHG soil moisture

product .................................................................................................................. 77

Figure 6.43 Landsat 8 satellite.................................................................................................. 80

Figure 6.44 Landsat 8 image footprint ..................................................................................... 81

Figure 6.45 Landsat images with different cloud-cover conditions ......................................... 82

Figure 6.46 Albedo map generated in ILWIS .......................................................................... 85

Figure 6.47 NDVI map generated in ILWIS ............................................................................ 86

xiii

Figure 6.48 Surface emissivity map generated in ILWIS ........................................................ 87

Figure 6.49 Land surface temperature map generated in ILWIS ............................................. 88

Figure 6.50 DEM map used in ILWIS as an input in SEBS .................................................... 88

Figure 6.51 The SEBS model in ILWIS ................................................................................... 89

Figure 6.52 Evaporative fraction map generated as an output of the SEBS model in

ILWIS.................................................................................................................... 90

Figure 6.53 Relative evaporation map generated as an output of the SEBS model in

ILWIS.................................................................................................................... 90

Figure 6.54 Catchment VI shapefile overlaid onto relative evaporation map .......................... 91

Figure 6.55 The relative evaporation map of Catchment VI .................................................... 91

Figure 6.56 Processes used to obtain relative evaporation from Landsat 8 data ...................... 93

Figure 7.1 Daily rainfall measured at Catchment VI ............................................................. 95

Figure 7.2 Daily in-situ and CRP soil moisture estimates for Catchment VI ........................ 96

Figure 7.3 Scatterplot of In-situ soil moisture estimates against CRP estimates ................... 97

Figure 7.4 Residual graph of In-situ against CRP soil moisture estimates ............................ 98

Figure 7.5 Time series analysis of CRP and AMSR2 soil moisture estimates..................... 100

Figure 7.6 Comparison between summer and winter images of AMSR2 soil moisture

estimates .............................................................................................................. 101

Figure 7.7 Scatterplot of CRP against AMSR2 soil moisture estimates .............................. 102

Figure 7.8 Residual graph of CRP against AMSR2 soil moisture estimates ....................... 103

Figure 7.9 A day with a descending and a day with an ascending value for the AMSR2

soil moisture product ........................................................................................... 104

xiv

Figure 7.10 A day with both an ascending and descending value, and a day with no

value for the AMSR2 soil moisture product ....................................................... 105

Figure 7.11 Time series analysis of CRP and SMOS soil moisture estimates for

Catchment VI ...................................................................................................... 106

Figure 7.12 Ascending SMOS image in summer (17 December 2014) ................................. 107

Figure 7.13 Descending SMOS image in summer (17 December 2014) ............................... 107

Figure 7.14 Ascending SMOS image in winter (15 August 2014) ........................................ 108

Figure 7.15 Descending SMOS image in winter (15 August 2014) ....................................... 108

Figure 7.16 Scatterplot of CRP against SMOS soil moisture estimates................................. 109

Figure 7.17 Residual graph of CRP against SMOS soil moisture estimates .......................... 109

Figure 7.18 SMOS missing data within band ......................................................................... 110

Figure 7.19 AMSR2, SMOS and CRP soil moisture estimates against time ......................... 111

Figure 7.20 Three-day averaged soil moisture estimates ....................................................... 112

Figure 7.21 Time series analysis of SAHG and CRP soil moisture estimation ..................... 114

Figure 7.22 SAHG daily soil moisture (summer) .................................................................. 115

Figure 7.23 SAHG daily soil moisture (winter) ..................................................................... 116

Figure 7.24 Scatter graph of CRP against SAHG soil moisture estimates ............................. 117

Figure 7.25 A residual graph of CRP against SAHG was plotted against time ..................... 118

Figure 7.26 A range of different relative evaporation images for Catchment VI .................. 121

Figure 7.27 A range of different evaporative fraction images ............................................... 122

Figure 7.28 Time series of CRP estimates and soil moisture back-calculated from the

SEBS model ........................................................................................................ 123

xv

Figure 7.29 Scatter graphs of CRP against the SEBS Model estimates a) CRP against Su

et al. (2003) and b) CRP against Scott et al. (2003). .......................................... 124

Figure 7.30 Residual graph of CRP against Su et al. (2003) and Scott et al. (2003). ............ 125

Figure 7.31 Time series of the CRP soil moisture estimates and soil moisture from the

Scott et al. (2003) method using the evaporative fraction from SEBS and

the eddy covariance method. ............................................................................... 127

Figure 7.32 Number of observation days per month that data was available for each

product ................................................................................................................ 129

xvi

LIST OF SYMBOLS

q p = Pore water content (g/g)

rbd = Dry soil bulk density (g/cm3)

qLW = Lattice water content (g/g)

qSOCeq = Soil organic carbon water content (g/g)

rv = Absolute humidity of air (g/m3)

AL = Band specific additive rescaling factor

Ap = Band specific additive rescaling factor

BWE = Biomass water equivalent (mm)

CI = High-energy intensity correction factor

CP = Pressure correction factor

CS = Geomagnetic latitude scaling factor

CWV = Water vapour correction factor

d = Earth sun distance

E = Evaporation (mm)

eo = Actual vapor pressure at surface (Pa)

eso = Saturated vapor pressure at surface (Pa)

Go = Soil heat flux energy (Wm-2)

H = Sensible heat flux energy (Wm-2)

Hdry = Sensible heat flux at the dry limit (Wm-2)

xvii

Hwet = Sensible heat flux at the wet limit (Wm-2)

Io = Irrigation (mm)

Lλ = TOA spectral radiance (Watts/ (m2*srad*µm))

ML = Band specific multiplicative rescaling factor

Mp = Band specific multiplicative reflectance factor

Ms = Mass of soil (g)

Mt = Mass of total (g)

Mvap = Molar mass of water vapor (= 18.01528 g/mol = 0.01801528 kg/mol)

N = Corrected neutron counts (counts/hour)

N’ = Raw moderated neutron counts (counts/hour)

Ɵ = Volumetric soil moisture (m3.m-3)

Ɵr = Residual volumetric soil moisture (m3.m-3)

Ɵsat = Saturated volumetric soil moisture (m3.m-3)

P = Pressure (mb)

Po = Precipitation (mm)

Ps = Particle density (g/cm3)

pλ = TOA planetary reflectance

pλ’ = TOA planetary reflectance (without solar angle correction)

Qcal = Quantized and calibrated standard product pixel value (DN)

R = Universal gas constant (= 8.31432 J/mol/K)

RH = Relative humidity (%)

xviii

Rn = Net radiation (Wm-2)

Rvap = Gas constant for water vapor (J/K/kg)

T = Air temperature (oC)

t = Time (s)

TC = Soil total carbon (g/g)

Ts = Land surface temperature (K)

Z = Vertical distance (m)

α = Land surface albedo (dimensionless)

εo = Surface emissivity (dimensionless)

ΘSE = Local sun elevation (degrees)

λE = Latent heat flux energy (Wm-2)

λEdry = The latent heat at the dry limit (Wm-2)

λEwet = The latent heat at the wet limit (Wm-2)

Λr = Relative evaporation (dimensionless)

xix

LIST OF ABBREVIATIONS

AMSR2 : Advanced Microwave Scanning Radiometer

CRP : Cosmic Ray Probe

ESA : European Space Agency

ILWIS : Integrated Land and Water Information System

LST : Land Surface Temperature

NDVI : Normalized Difference Vegetation Index

PyTOPKAPI : Python Topographic Kinematic Approximation and Integration

SAHG : Satellite Applications and Hydrology Group

SEBAL : Surface Energy Balance Algorithm for Land

SEBS : Surface Energy Balance System

SMOS : Soil Moisture and Ocean Salinity

TDR : Time Domain Reflectometry

TOPKAPI : Topographic Kinematic Approximation and Integration

VWC : Volumetric Water Content

1

1. INTRODUCTION

There has been an incessant need to monitor and measure the various parameters in land

surface hydrology, in order to deepen the understanding of hydrological processes, their

importance in the hydrological cycle and their interactions between each other (Jackson et al.,

2010; Fang and Lakshmi, 2014). Soil moisture is an important parameter in the hydrological

cycle and greatly impacts a variety of applications, including agricultural management,

climate and weather applications, flood and drought prediction and groundwater recharge.

Although soil moisture does not directly contribute to the land surface energy balance, the

properties associated with soil moisture have a strong influence on net radiation, and the

partitioning between latent and sensible heat fluxes (Wang and Li, 2011). Soil moisture is a

significant variable in the earth’s water cycle, even though it holds a small percentage of the

total global water budget (Guillem, 2010). Weather and climate applications rely greatly on

soil moisture, consequently the predication accuracy of global weather forecast models and

climate applications may increase, with improved estimates of soil moisture data (Guillem,

2010). Soil moisture is a key component in crop growth and is the prime regulator of a

catchment’s response to rainfall, as it partitions rainfall into infiltration and surface runoff

(Vischel et al., 2008; Zhao and Li, 2013). Due to the impacts of global change, numerous

cycles, such as the hydrological cycle, have been pushed past their thresholds. Therefore,

monitoring the soil moisture status of an area, such as a catchment, can assist in mitigating

the negative effects of extreme environmental processes or phenomena, such as droughts and

floods (Guillem, 2010).

Soil moisture is a difficult parameter to continuously monitor and measure at a catchment

scale due to its heterogeneous characteristics. It varies both spatially and temporally and is

thus a dynamic resource, which never remains constant. Currently, there are three methods of

estimating soil moisture. These include: (a) ground-based (in-situ) measurements, which are

carried out using field instruments; (b) remote sensing based methods, which use specialised

sensors on satellites and aircrafts and (c) land surface models, which use meteorological data

as inputs, at a predefined spatial resolution (Albergel et al., 2012; Mecklenburg et al., 2013).

Inherently each of these methods possess their respective advantages and limitations,

constraining their effectiveness for hydrological applications (Ni-Meister, 2005).

2

In-situ measurements of soil moisture are the conventional methods used by several

disciplines. The point measurements obtained cannot adequately represent the spatial

characteristics of soil moisture. However, these point measurements play a key role in a

variety of large-scale applications and are invaluable as both calibration and validation data

(Gruber et al., 2013). In the past few years, the Cosmic Ray Probe (CRP) method, which is an

in-situ instrument, has been developed and implemented in numerous countries across the

globe (Kohli et al., 2015). The CRP technique obtains the area-averaged soil moisture, at an

intermediate scale, by observing and measuring the cosmic ray neutrons above the soil

surface (Zreda et al., 2012; Franz et al., 2013).

Remote sensing can be described as the technique of obtaining information about an object or

phenomena, with the use of special instrumentation from a distance, without making physical

contact (Engman, 1991; Wagner, 2008; Mekonnen, 2009; Lakshmi, 2013). Remote sensing is

extensively used today by several disciplines to observe and measure various environmental

parameters. Hydrology is one such discipline, which uses remote sensing to measure

parameters, such as rainfall, evaporation and surface temperature (Fang and Lakshmi, 2014).

There are several remote sensing soil moisture estimating techniques, which use gamma, near

infrared, microwave and thermal radiation. The passive microwave radiation technique is the

most commonly used technique, as it has several advantages over other techniques. The

passive microwave technique observes and measures the difference in dielectric constants of

dry soil and water to estimate soil moisture (Jackson et al., 2010; Lakshmi, 2013). Currently,

the microwave remote sensing soil moisture products are available on a global scale;

however, these products are limited by their coarse spatial resolution, which is inadequate for

hydrological applications (Shin and Mohanty, 2013; Song et al., 2013). Therefore,

downscaling techniques have been researched and developed, to obtain soil moisture at a

finer spatial resolution. (Lakshmi, 2013).

Models have been used to obtain estimates of soil moisture at an intermediate scale. A model

implies a simplification of reality and aims to promote understanding, whilst simplifying and

mimicking the real system. Soil moisture is an important variable in both the land surface

energy and water balance and can be estimated from surface energy balance models, such as

the Surface Energy Balance System (SEBS) and land surface water models, such as the

Topographic Kinematic Approximation and Integration Model in Python (PyTOPKAPI).

3

The objective of this study is to review literature pertaining to the current methods of soil

moisture estimation, which include in-situ, remote sensing and modelling. This is required to

broaden the understanding of this field of study. This will aid in identifying the

commonalities and gaps in the literature, as well as the key methods used and their limitations

in the estimation of soil moisture. The key focal aspects of the literature review are to

evaluate the CRP for use as a validation method. In addition, the current methods of remote

sensing soil moisture products will be reviewed and the modeled soil moisture estimates will

be analysed. The aim of the study is to test and evaluate the suitability of the CRP to provide

spatial estimates of soil moisture in Cathedral Peak Catchment VI and use these estimates to

validate satellite-based soil moisture estimates. Satellite-based soil moisture products

comprise of remote sensing soil moisture products and models, which use remote sensing

data as inputs.

The Research questions include:

i. How suitable is the CRP method in providing spatial estimates of soil moisture in

Cathedral Peak Catchment VI?

ii. How well do remote sensing soil moisture products compare to the CRP estimates?

iii. How well do modelled soil moisture products compare to the CRP estimates?

The dissertation is divided into eight chapters, starting with the introduction in chapter one.

The literature review is not a single chapter but is encompassed in chapters two, three, four

and five. Chapter two outlines the in-situ methods of soil moisture measurement, which

consists of conventional in-situ methods, as well as the CRP. Chapter three outlines the

remote sensing of soil moisture and focuses on the current global soil moisture products.

Chapter four outlines the use of modelling to obtain soil moisture estimates. Chapter five is

the literature evaluation, which looks at the general themes in soil moisture measurements

and evaluates the gaps in the literature reviewed. Chapter five is an important chapter, as it

summarizes the evaluated literature and sets the foundation for the methodology, which is

presented in chapter six. The methodology is presented in chapter six, which outlines the

study area, as well as details the various procedures and techniques that were used in this

research study. Chapter seven contains the results and discussion of the research study. The

conclusions, limitations and recommendations of the study are discussed in chapter eight.

4

2. IN-SITU METHODS OF SOIL MOISTURE MEASUREMENT

The following section is the first component of the literature review. The remaining

components of the literature review will be covered in the third and fourth chapters. A

literature evaluation will be presented in chapter five.

In-situ soil moisture techniques are the conventional methods of soil moisture estimation.

There are numerous in-situ techniques, both direct and indirect, which have been used to

determine soil moisture. In recent years, the use of the CRP to measure soil moisture has

gradually attracted attention and has been implemented in a number of countries across the

world (Villarreyes et al., 2013). The ensuing section will provide an analysis of the

conventional methods of soil moisture estimation, which will be followed by a detailed

review of the CRP.

2.1 Conventional Methods of Soil Moisture Estimation

In-situ soil moisture measurements have played a key role for a variety of large-scale

applications and have been invaluable as calibration and validation data for satellite-based

products, sensors and models (Gruber et al., 2013). There are several conventional in-situ

techniques used to estimate soil moisture, including the gravimetric method, neutron

scattering and Time Domain Reflectance (TDR) (Walker et al., 2004). Brief descriptions of

the aforementioned techniques are discussed below:

i. The gravimetric method is a direct method of estimating the soil moisture content.

This method consists of collecting soil samples at points at desired depths. The

collected soil samples are then weighed, before being oven-dried for 24-72 hours at

105 oC, until all the moisture has been driven out of the sample. The dry samples are

then re-weighed. The gravimetric soil moisture is then calculated by dividing the mass

of water by the mass of dry soil.

This method is commonly used as it has the advantage of being accurate, is

independent of soil type and is easily calculated (Hillel, 2008). However, it is a

destructive technique that is time and labour-consuming (Zazueta and Xin, 1994).

5

Due to its destructive nature, it does not allow for repetitive sampling at the same

point. The method also requires a laboratory equipped with an oven and the time of

the sample collection to the time that the samples are weighed has to be minimal, to

avoid the loss of moisture from the soil (Hillel, 2008).

ii. The neutron scattering method is an indirect in-situ measurement, which has been

used since the 1950’s (Jones et al., 2002). The neutron probe uses an active source of

fast neutrons, which pass through the soil and collide with the soil media. These fast

neutrons thus slow down with each collision and become slow neutrons. The majority

of these fast neutrons collide with hydrogen in the form of water in the soil media.

Thus, the more water present in the soil media, the more hydrogen and the more

collisions occur. The measured slow neutrons are therefore proportional to the soil

water.

iii. According to Zazueta and Xin (1994), this method has the advantage of being

accurate (once calibrated) and able to measure soil profiles, with the measurement

being directly related to soil moisture. However, the method is limited because of its

cost, radiation hazard, the skills required and the difficulty to obtain accurate

measurements near the surface, because it is time-consuming and its invasive nature,

as access tubes have to be placed vertically into the soil profile.

iv. The TDR method is a common indirect in-situ method. According to Jones et al.

(2002), there are different variations of TDR probes, but in general, the probe consists

of two to three steel rods, which can be placed vertically or horizontally into the soil.

The probe or logger sends out high frequency pulses in the Giga Hertz (GHz) range,

which have a varying travel time along the wave guides(stainless steel rods), due to

the dielectric constant of the medium in which the probe is placed. The dielectric

constant is mostly dependent on the water content of the soil. The TDR probe is

insensitive to soil composition and texture, due to the large disparity of dielectric

constants, but needs calibration with soils with a high electrical conductivity (Jones et

al., 2002).

The TDR method is a commonly used and accurate method, due to its minimal

calibration requirements, easily obtainable measurements, low cost compared to other

6

methods, and its ability to provide non-destructive continuous measurements (Jones et

al., 2002; Walker et al., 2004). The TDR method is limited, as a calibration procedure

is required, the depth measured is dependent on the length of the steel rod used and

errors may occur from soil voids around the probes measurement area (Zazueta and

Xin, 1994; Walker et al., 2004) .

The major limitation of all conventional in-situ methods is that they provide point

measurements, which do not account for the spatial characteristics of soil moisture (Qin et

al., 2013). Therefore, no single point measurement can be entirely representative of larger

areas, because of the heterogeneity that exists in soil properties, topography and weather

(Gruber et al., 2013). In order to overcome this limitation, dense soil moisture networks can

be set up. However, the high costs of operation and maintenance make the setup of the

network financially unfeasible (Gruber et al., 2013).

2.2 Cosmic Ray Probe

Understanding the variability of soil moisture, is of great interest for various disciplines

(Villarreyes et al., 2013). The CRP is a relatively new technique, which has the capability of

providing soil moisture data for large-scale studies. It is the only in-situ technique that can

obtain the average soil moisture content over hundreds of meters, something that would

require a dense in-situ point measurement network (Zreda et al., 2012; Ochsner et al., 2013).

The CRP method can play a vital role in the calibration and validation of satellite-based soil

moisture retrievals and land surface models (Villarreyes et al., 2013).

As shown in Figure 2.1, the CRP system comprises of neutron counters (moderated and bare

tubes), a data logger which includes barometric pressure, humidity and temperature sensors,

and a telemetry system with antenna to connect to an iridium satellite, and a battery and solar

panel for powering the system.

7

Figure 2.1 CRP system in Cathedral Peak Catchment VI

Production of cosmic ray neutrons

Cosmic ray neutrons originate in space, where they are produced by the blast waves of

exploding stars. Shards of atoms are accelerated and energized, until they eventually reach a

high enough speed to break away and escape to the galaxy. Some of these high energy

particle flows in space reach the earth’s atmosphere, where they are affected by the earth’s

magnetic field (Jiao et al., 2014). The high energy particles are captured into the earth’s

atmosphere and collide with atmospheric nuclei to initiate a cascade of secondary cosmic

rays (Ochsner et al., 2013). Fast neutrons are created, as these secondary cosmic rays pass

through the atmosphere and then through the top few meters of the biosphere, hydrosphere

and lithosphere, (Desilets et al., 2010).

These fast neutrons undergo elastic collisions with nuclei present in the soil, thereby losing

energy (Desilets et al., 2010). Some of the fast neutrons are adsorbed by the soil during the

collision, whilst others will be scattered above the surface of the soil (Jiao et al., 2014). The

cosmic ray neutrons lose energy with each collision, therefore high energy neutrons become

fast neutrons (in the atmosphere), which further lose energy and become thermal neutrons (in

Satellite link

Neutron tubes

Data logger

Batteries

Pressure,

temperature

and

humidity

sensors

Solar panel

8

the soil). Due to fast neutrons being strongly moderated by hydrogen present in the

environment, their measured intensities relate to changes in soil moisture, as well as other

hydrogen sources at the earth’s surface (Zreda et al., 2008; Zreda et al., 2012; Franz et al.,

2013).

Moderation of neutrons

According to Ochsner et al. (2013) and Jiao et al. (2014) the moderation process of cosmic

ray neutrons depends on three factors:

i. The scattering probability or the elemental scattering cross-section;

ii. The logarithmic decrement of energy per collision; and

iii. The number of atoms of an element per unit mass of material, which is proportional to

the concentration of the element and to the inverse of its mass number.

The combination of the abovementioned factors, defines the neutron stopping power of a

material (Ochsner et al., 2013). Hydrogen, which is found mainly as water in the soil, plays

the most significant role in moderating cosmic ray neutrons in the soil. Hydrogen has an

extraordinary high stopping power, as the hydrogen atom has a high probability of scattering

a neutron, due to its fairly large elastic scattering cross-section (Jiao et al., 2014). Hydrogen

is the most efficient element, with regards to the decrement of energy per collision and has a

low atomic mass and makes up a substantial portion of all the atoms in many soils, due to the

presence of water in the soil (Jiao et al., 2014). The presence of water within the soil pores

plays an important and central role in moderating the concentration of cosmic ray neutrons

above the soil surface (Desilets and Zreda, 2013).

Cosmic ray probe measurements

The fast neutrons that are produced in the air and soil travel in all directions between the air

and soil, thus creating an equilibrium concentration of neutrons. This equilibrium

concentration is shifted due to changes (addition or subtraction) in the hydrogen content of

the media. The soil moisture content is estimated by the concentration of low energy cosmic

ray neutrons, which are generated within the soil and moderated predominantly by hydrogen,

9

before being diffused back into the atmosphere (Zreda et al., 2008). The soil moisture content

can therefore be inferred directly from these neutron fluxes (Desilets et al., 2010).

Dry soils are highly emissive, so that neutrons are more efficiently removed from the soil

(Zreda et al., 2008). This results in more neutrons escaping to the surface of a dry soil, which

will result in a higher concentration of neutrons above the soil surface, see Figure 2.2 (Franz

et al., 2012b).

Figure 2.2 Difference in neutron concentration according to soil moisture content

(Franz et al., 2012b).

The CRP system consists of two sensors. The moderated sensor (encased in perspex)

measures the fast neutrons, which are attributed to the soil moisture, whilst the bare tube

measures the slow neutrons, which are attributed to the water above the soil surface (biomass

and snow). The fast neutron intensity above the soil surface is inversely proportional to the

amount of soil moisture in the soil surface (Kohli et al., 2015).

Measurement footprint and depth of the cosmic ray probe

The CRP senses all hydrogen present within the distance that fast neutrons can travel in air,

water, soil and other materials near the earth’s surface. Thus, the measurement distance varies

according to the density and chemical composition of the material (Ochsner et al., 2013). The

footprint (measurement area) of the CRP, is defined as the area around the probe from which

86% of the counted neutrons arise (Jiao et al., 2014). The footprint is primarily associated

Dry Soil Wet Soil

10

with the chemical and physical properties of the air and is inversely proportional to the air

density (Jiao et al., 2014).

The radial footprint of the CRP is reliant on the neutron’s ability to travel hundreds of meters

from their source, due to the neutrons scattering in the air (Kohli et al., 2015). Hence the

scattering properties of air significantly affect the diameter of the footprint (Jiao et al., 2014).

When the CRP is placed in a static position a few meters above the ground, it has a radial

footprint of 670 meters in diameter at sea level (Zreda et al., 2008). The radial footprint size

is inversely proportional to the air density. At high altitudes, the air pressure is lower, which

results in a larger footprint. The measurement diameter slightly decreases with increasing soil

moisture and with increasing atmospheric water content (Jiao et al., 2014).

The effective depth is the thickness at which 86% of counted neutrons arise, which depends

strongly on the soil moisture content (Ochsner et al., 2013). The measurement depth is

inversely proportional to the soil moisture content. A measurement depth of 0.72 m is

obtained in dry soil and a depth of 0.12 m is obtained in wet soil, so that the effective

measurement depth decreases with increasing hydrogen. (Zreda et al., 2008). The decrease in

measurement depth according to an increase in soil moisture is nonlinear and the

measurement depth is independent of the air density.

Due to the fact that neutrons react with any source of hydrogen near the earth’s surface, the

measured neutron intensity represents the total reservoir of neutrons present within the

probe’s sensing distance (Ochsner et al., 2013). This includes snow, lattice water, water in

soil organic matter, water in vegetation and atmospheric water vapour (Ochsner et al., 2013).

The measurement footprint and depth is shown in Figure 2.3.

Figure 2.3 Measurement area and depth of CRP (Ochsner et al., 2013)

11

The CRP can be used either in a fixed position or in a moving vehicle. The fixed position is

used to obtain continuous measurements of an area, whilst the roving method can be used for

mapping large areas (Dutta and D'este, 2013). The technique functions, as the neutron fluxes

are a great proxy for land surface water (Desilets et al., 2010). Along with the neutron count

rate, the CRP also measures the internal temperature, relative humidity within the logger

enclosure and external barometric pressure (Franz et al., 2013).

Advantages of the cosmic ray probe method

The CRP method has several advantages (Zreda et al., 2008; Desilets et al., 2010; Franz et

al., 2012b; Franz et al., 2012a; Zreda et al., 2012; Desilets and Zreda, 2013; Franz et al.,

2013; Jiao et al., 2014), including:

Table 2.1 Advantages of the CRP

Advantages

The method is passive and non-contact (non-invasive)

The system is easily automated

The CRP is portable

It has minimal power requirements Applications not limited to soil moisture (can be used to estimate above-ground biomass and snow depth)

The measurement footprint is at an intermediate scale

The measurement depth ranges from 0.12 m to 0.72 m

The method allows for continuous measurements to be obtained

The instrument is easily installed above-ground

The CRP provides excellent data sets

The CRP requires low data processing and transmission

The method is less affected by the presence of vegetation

The method is insensitive to soil texture, bulk density or surface roughness

12

The CRP is a fairly recent instrument used to obtain soil moisture estimates and seems to

have no major limitation associated with it. However, the measurement depth and the

calibration procedure could be potential limitations of the technique. The measurement depth

is not set by the user, but is dependent on the soil moisture status of the measurement area.

The measurement depth of the CRP is generally between 0.12 and 0.72 meters, therefore the

CRP technique is limited to measuring soil moisture between this depth range. The

calibration procedure is time and labour consuming and it needs to be done correctly in order

to provide reliable soil moisture estimates.

The CRP calibration procedure is performed by obtaining corresponding measurements of

area-averaged soil moisture and neutron intensity. The area-averaged soil moisture is

obtained from many ground-based point measurements, by collecting several soil samples

within the CRP measurement area and determining the average gravimetric soil moisture per

calibration. The measured neutron intensities need to be adjusted and corrected for variations

in location, incoming high-energy particles, atmospheric pressure, absolute humidity and

changes in biomass (Franz et al., 2015).

It is recommended that a calibration procedure is carried out for both the dry season and the

wet season (Dutta and D'este, 2013). Representative soil samples of the measurement area are

required to be analysed, to correct the calibration function for lattice water and water in

organic matter (Ochsner et al., 2013).

The potential applications of the CRP make it appealing to scientists in various fields, such as

agricultural and ecological monitoring, streamflow forecasting, climate science, drought and

flood forecasting, as well as slope stability (Desilets et al., 2010). It should also be noted that

the discipline of remote sensing can benefit greatly from this innovative technology, by using

CRP measurements for both the calibration and validation of sensors and data products, as it

overcomes the spatial limitations of conventional in-situ soil moisture estimates (Desilets et

al., 2010). The use of the CRP across different continents results in measurement technique

consistency, which would reduce the uncertainties related to in-situ measurements that use a

variety of probe types/methods (Kim et al., 2015).

There have been a limited number of studies using the CRP, as the technique is fairly new. A

summary of these studies is presented below:

13

i. Zreda et al. (2008) conducted a study in Arizona, USA, on the use of cosmic ray

neutrons to estimate soil moisture at an intermediate spatial scale. The study

concluded that the CRP permits high resolution long-term monitoring and the large

radial measurement footprint makes the CRP method suitable for climate and weather

applications, as well as for satellite sensor calibration.

ii. Desilets et al. (2010) looked at implementing the CRP in two ways. The CRP was

used to provide stationary measurements in south-eastern Arizona and was used to

provide roving measurements in Hawaii. The study concluded that both methods can

be implemented with confidence.

iii. Desilets and Zreda (2013) investigated the lateral footprint of the CRP, using a

combination of neutron transport simulations and the diffusion theory. The study

concluded that the measurement footprint is linearly proportional to the sensor height

above the ground and inversely proportional to the air density.

iv. Hawdon et al. (2014) conducted a study, which looked at the set-up and field

calibration of the CRP at nine locations across Australia. The nine locations had

contrasting soil types, land covers and climates. The results of the study indicated that

the generalized calibration function is suitable for relating neutron counts to soil

moisture and that it holds for all soil types, except for very sandy soils with lower

water contents.

14

3. REMOTE SENSING OF SOIL MOISTURE

Remote sensing is currently used to measure various parameters of the hydrological cycle,

such as rainfall, evaporation and soil moisture. Remote sensing is seen as a promising

technique for soil moisture estimation, as it aims to address the spatial and temporal

variability of soil moisture (Zhao and Li, 2013). It has the advantages of being able to

measure large scale soil moisture, the data obtained has periodic data updates, its ability to

monitor soil moisture in remote areas, its usability and its cost (Sabins, 2007; Mekonnen,

2009; Lakshmi, 2013). The major limitation to the implementation of remote sensing in

critical hydrological application is its coarse resolution. The following section gives an

overview of the common remote sensing of soil moisture products. This is followed by an

analysis and evaluation of the current downscaling techniques that have been developed to

obtain soil moisture at a finer spatial resolution.

3.1 Overview of Remote Sensing of Soil Moisture

There are several remote sensing techniques that have been developed and applied, which

include gamma radiation, thermal infrared, near infrared and microwave radiation techniques

(Albergel et al., 2012). Each technique measures a different land surface quantity, uses a

different range of the electromagnetic spectrum and has its own unique advantages and

limitations (Mekonnen, 2009). From past research studies, it is evident that the microwave

radiation technique, which consists of both active and passive methods, can be considered as

the most promising technique for the remote sensing of soil moisture. This is due to its

advantages over the other techniques, such as its all-weather capability, large spatial

coverage, temporal resolution, measurement depth and vegetative penetration (Wagner, 2008;

Wang and Qu, 2009; Guillem, 2010).

Microwave radiation remote sensing observes the large contrast in the dielectric properties of

soil particles and water, as well as the dielectric constant increases, as the soil moisture

increases (Mekonnen, 2009; Wang and Qu, 2009). Remote sensing does not measure the soil

moisture content directly, therefore mathematical models that describe the association

between the measured signal and the subsequent soil moisture need to be derived (Wang and

Qu, 2009). Over the past few decades, active and passive microwave remote sensing has

provided the unique ability to obtain estimates of soil moisture on a global scale (Brocca et

15

al., 2013). The L-band range (1 to 10 GHz) is preferably used, as higher frequencies are more

affected by perturbation factors, such as vegetation cover and atmospheric effects (Albergel

et al., 2012). There have been many research studies on microwave remote sensing products

over the past few decades. These products include Soil Moisture and Ocean Salinity (SMOS),

the Advanced Microwave Scanning Radiometer (AMSR2), European Remote Sensing

Satellite (ERS-1/2) and Advanced Scatterometer (ASCAT) (Dorigo et al., 2011).

The remote sensing of soil moisture has advanced and the product user’s trust in the remote

sensing data has increased, as there has been continuous improvements to sensors and the

algorithms used to estimate soil moisture (Brocca et al., 2013). The launch of the SMOS

satellite, which was the first satellite radiometer dedicated to measure soil moisture over land,

emphasized the increased need for the measurement of soil moisture. This is further

highlighted by the anticipated launch of the Soil Moisture Active Passive (SMAP) satellite in

2014 (Jackson et al., 2010; Lakshmi, 2013; Song et al., 2013; Fang and Lakshmi, 2014).

Currently, there are several remote sensing soil moisture products. The two most frequently

utilized products that appear in recent research literature, are the AMSR/AMSR-E

(predecessors of the AMSR2) and SMOS soil moisture products (Brocca et al., 2013). These

two products will be detailed in the subsequent sections.

3.2 AMSR2

The global change observation mission (GCOM) is a project for the global long-term

monitoring of the earth’s environment. The GCOM consists of two satellites, namely, the

GCOM-W and GCOM-C. The GCOM-W1 satellite was launched on the 18th of May 2012

and is equipped with the AMSR2 sensor (Kim et al., 2015). The AMSR2 sensor on the

GCOM-W1 satellite, as seen in Figure 3.1, is the successor to the AMSR sensor on-board the

ADEOS-II satellite, which has been operational from December 2002 to October 2003 and

the AMSR-E sensor, on-board the AQUA satellite, which has been operational from May

2002 to October 2011 (JAXA, 2013a). The basic performance of AMSR2 will be similar to

that of AMSR-E, based on the minimum requirement of data continuity (Imaoka et al., 2010).

16

Figure 3.1 AMSR2 sensor on-board the GCOM-W1 satellite

The AMSR2 sensor is a passive microwave scanning radiometer that makes observations at

multiple polarizations (both horizontal and vertical) and seven frequency bands, ranging from

6.9 GHz to 89 GHz (Kim et al., 2015). It obtains multi-band observations from the C, X and

K bands (Lu et al., 2014). The waves observed are weak and are radiated from the

atmosphere and the earth’s surface. It orbits 700 km above the earth’s surface and has a swath

of 1450 km. Its antenna rotates once every one and a half seconds, which creates a conical

scan pattern, with a scanning interval of 10 km. The scanning method is capable of making

one day (ascending) and one night (descending) observation, covering more than 99% of the

earth’s surface in two days (Lu et al., 2014).

The AMSR2 sensor estimates several of the parameters, which are predominantly linked to

the energy and water cycles, such as precipitation, water vapour, sea-surface temperature, soil

moisture and snow depth. The AMSR2 sensor is more advanced than its predecessors

(AMSR and AMSR-E). It has a larger main reflector with a 2.0 m antenna diameter, which

results in a finer spatial resolution, the addition of a new 7.3 GHz channel to eliminate

electro-magnetic wave interference, an upgraded calibration system and it has intensive

sunlight shielding, to avoid calibration uncertainties (Imaoka et al., 2010; JAXA, 2013a; Kim

et al., 2015)

17

According to JAXA (2013b), the AMSR2 Level Two soil moisture product uses auxiliary

data in the form of global vegetation coverage ratios. It also uses AMSR2 Level One

products, which include 10.6 GHz brightness temperatures (horizontal and vertical

polarisations), 36.5 GHz brightness temperatures (horizontal polarization), latitude and

longitude and observation date.

The AMSR2 Level Two algorithm uses a simple radiative transfer model. The satellite

observed brightness temperature (TB) at the 10.7 GHZ and 36.5 GHz channels, measures the

natural microwave emissions from the land surface (Kim et al., 2015). The algorithm

assumes that the soil temperature is equal to the vegetation canopy temperature and a

constant value of 293 K is used throughout the year (Kim et al., 2015). The AMSR2 soil

moisture algorithm assesses the vegetation effects on microwave emissions paths. This is

important, as vegetation holds a large amount of moisture and thus needs to be evaluated in

order to estimate soil moisture from satellite data. The soil moisture algorithm uses the

polarization index and the index of soil wetness, obtained from TB, to simultaneously

estimate soil moisture and vegetation moisture. With regards to vegetation, the algorithm uses

a look-up table method, where a linear relationship between vegetation water content and

optical depth is applied (Kim et al., 2015). A dielectric mixing model is used in the algorithm

to convert the dielectric constant into soil moisture content (Kim et al., 2015).

3.3 SMOS

The SMOS satellite system, which is illustrated in Figure 3.2, was launched in November

2009 and consists of a space-borne passive microwave L-band (1.4 GHz) interferometric

radiometer (Albergel et al., 2012). The overall aim of SMOS is to provide global surface soil

moisture maps with an accuracy of 4% at a 35 – 50 km spatial resolution (Parrens et al.,

2012). The SMOS satellite has the ability to obtain measurements at multiple angles, which

allows for the retrieval of additional parameters (Kerr et al., 2012). The SMOS system is a Y-

shaped instrument, which consists of 69 antennae. The antennae are equally spaced along

three arms and view the surface of the earth either through full, or two polarized radiances, in

order to provide a full image (Gruhier et al., 2011).

18

Figure 3.2 SMOS satellite

SMOS operates by measuring the phase difference of radiation from various incident angles,

so that the earth’s surface is frequently viewed at different angles and polarizations (Fang and

Lakshmi, 2014). The SMOS soil moisture product has an average spatial resolution of 40 km;

however, this varies from 30-50 km due to, the target area being viewed from multiple angles

(Kerr et al., 2010). SMOS has a sun-synchronous orbit at an altitude of 757 km,

consequently, the entire earth is covered at least twice in three days, with orbit overpasses at

6:00 (ascending) and 18:00 (descending) local standard time (Leroux et al., 2013; Qin et al.,

2013).

L-band radiometry has a high potential for the estimation of surface parameters and has been

recognised as the most promising technique to monitor soil moisture over land surfaces and at

a global scale (Kerr et al., 2012; Parrens et al., 2012). The L-band is the optimum wavelength

range to observe soil moisture, as higher frequencies are more affected by perturbing factors,

including atmospheric effects and vegetative cover (Parrens et al., 2012; Mecklenburg et al.,

2013). The microwave signal at L-band frequency is primarily driven by surface soil moisture

(top 5 cm), effective surface temperature and vegetative opacity, with smaller effects from

the atmosphere and surface characteristics, such as bulk density, soil texture and surface

roughness (Kerr et al., 2012; Leroux et al., 2013). The soil moisture retrieval algorithm needs

to account for all these effects (Kerr et al., 2012).

19

The SMOS Level Two retrieval algorithm is physically-based and complex. It uses the

polarizations and multi-angular information derived by the SMOS satellite and involves the

use of decision trees and direct model inversions (Parrens et al., 2012). The retrieval

algorithm was designed to be easily improved and uses SMOS Level 1C data as input and

uses two types of auxiliary data, namely, static data (soil texture) and dynamic data

(temperature) (Kerr et al., 2012). SMOS provides multi-angular microwave polarimetric

brightness temperature (TB), from which a surface soil moisture product is derived (Parrens et

al., 2012).

The L-band microwave emission of the biosphere model, which simulates the microwave

emission at L-Band range from the soil-vegetation layer is a key component of the retrieval

algorithm (Parrens et al., 2012). This is important, as the major difficulty in the estimation of

soil moisture, using microwave radiometry is due to the presence of dense overlying

vegetation, which attenuates the soil emission and adds its own emission (Parrens et al.,

2012).

The retrieval algorithm is based on an iterative approach. The aim of the algorithm is to

reduce a cost function, the focal constituent of which is the sum of the squared weighted

difference between measured TB and Modelled TB for a collection of incidence angles (Kerr

et al., 2012). In order to reduce this cost function, the most suitable set of parameters needs to

be determined and used to drive the TB model.

Current remote sensing soil moisture products have a spatial pixel grid between 10 to 50 km.

The major issues in the calibration and validation procedure, when using in-situ point

measurements, are the vertical and horizontal scaling issues. The vertical scaling issues occur

when remote sensing soil moisture (top 0.05 m) is calibrated and validated against in-situ soil

moisture measurements (0.00 to 2.00 m) (Jackson et al., 2010). Therefore, the measurements

are at different depths, which is a problem, as soil moisture varies with depth. The horizontal

scaling issues occur when a point measurement is used to validate a remote sensing area-

averaged value. The assumption that the point is representative of a large area can be

conceded to be incorrect, due to the spatial variability of soil moisture (Gruber et al., 2013).

Therefore, there needs to be a shift to area-averaged in-situ methods, to validate and calibrate

remote sensing data.

20

There have been numerous studies on passive microwave remote sensing for soil moisture

estimation. The AMSR-E product is a frequently-used product in past studies. Research

studies that have evaluated surface soil moisture derived from AMSR–E, have shown

promising results (Koike et al., 2004; Draper et al., 2009; Zhang et al., 2011; Qin et al.,

2013). Previous literature has mentioned the launch of the SMOS satellite and how it

emphasizes the value of soil moisture estimation in the science field. SMOS data have been

used in studies which have shown its potential for accurately obtaining soil moisture

measurements (Gruhier et al., 2011; Merlin et al., 2012; Wagner et al., 2012; Mecklenburg et

al., 2013; Qin et al., 2013). Over the past few decades, there have been many studies

involving both the calibration and validation of remote sensing soil moisture products to in-

situ soil moisture measurements (Jackson et al., 2010; Brocca et al., 2011; Fang and

Lakshmi, 2014). The following are some of the remote sensing based soil moisture product

studies:

i. The AMSR2 soil moisture product has been used in a recent study by Lu et al. (2014),

which presented a revised soil moisture retrieval algorithm of AMSR2 and aimed at

improving the soil moisture estimates in dry regions. The study concluded that the

revised algorithm is effective in overcoming the over-estimation of soil moisture in

desert regions.

ii. A recent study by Kim et al. (2015) assessed two versions of the AMSR2 soil

moisture product. The AMSR2 product from the Japan Aerospace Exploration

Agency (JAXA) and Land Parameter Retrieval Model (LPRM) from the VU

University Amsterdam, were used and compared on a global scale against field

measurements. The study used CRP measurements from the COSMOS. It concluded

that the JAXA AMSR2 product has a relatively better performance under dry

conditions; however, the JAXA algorithm generally under-estimates the ground soil

moisture. It also indicated that there is a need to improve both the algorithms and that

a combined product could provide better estimates of soil moisture.

iii. A study by Leroux et al. (2013), compared the SMOS, AMSR-E and ASCAT soil

moisture products over four watersheds in the United States for the year 2010. The

study concluded that the SMOS product correlated the closest to ground

21

measurements, the AMSR-E product gives reasonable results in terms of correlation

and the ASCAT product was unstable.

iv. A study conducted by Albergel et al. (2012) evaluated the ASCAT and SMOS

products against in-situ soil moisture observations from over 200 stations across the

world for the year 2010. A similar study was conducted by Brocca et al. (2011),

which evaluated the ASCAT and AMSR-E satellite-based soil moisture products

around Europe. The main purpose was to evaluate the potential of different ASCAT

and AMSR-E products in obtaining reliable estimates of soil moisture. The study

concluded that the AMSR-E LPRM provided the best results.

3.4 Downscaling Techniques

Microwave remote sensing is extensively used to obtain regional soil moisture estimates,

although its application is greatly restricted by its coarse resolution (Zhao and Li, 2013).

There have been several research studies undertaken to downscale satellite-based soil

moisture estimates for use in hydrological applications (Song et al., 2013). Downscaling is

relevant, as the coarse resolution is at tens to hundreds of kilometres, whereas hydrological

processes occur at spatial scales of centimetres to kilometres (Shin and Mohanty, 2013).

Satellites have the capability to observe and monitor large areas. The subsequent

observational spatial resolution is dependent on the microwave frequency used, the

dimensions of the antenna and the height of the satellite above the earth’s surface (Piles et al.,

2011). There is a trade-off between the temporal and spatial satellite resolution; as the sensor

height decreases, the spatial resolution increases and the temporal resolution decreases

(Merlin et al., 2012).

Over the past decade, a number of downscaling techniques, with varying degrees of

complexity, have been researched and developed (Merlin et al., 2008). An extensive research

paper by Lakshmi (2013) outlines the key publications on current methods used to downscale

passive microwave remote sensing soil moisture products. These studies have reviewed

downscaling procedures, with the use of MODIS sensor-derived parameters, including

surface temperature, vegetation and other ground surface parameters.

22

The consistent theme, with regards to downscaling procedures, which is identified in these

key publications, is the combined use of passive microwave data with fine-scale optical data

(surface temperature and vegetative indexes). The overall aim of these downscaling

techniques are to provide soil moisture estimates at the same accuracy as the input remote

sensing soil moisture product, but at the spatial resolution of the optical data used. These key

downscaling publications are evaluated and summarized in Table 3.1 below.

Table 3.1 Evaluation of downscaling techniques Author Merlin et al Piles et al Merlin et al Merlin et al

Year 2012 2011 2010 2009

Region Yanco, South-eastern

Australia (2010)

Yanco, South-eastern

Australia (2010)

Yanco, South-eastern

Australia (2006)

Yanco, South-eastern

Australia (2006)

Input data MODIS (LST,

emissivity, NDVI and

Albedo)

MODIS VIS/IR data (

LST and NDVI)

MODIS (LST, red and

infrared reflectance and

NDVI)

Wind speed, MODIS

data (surface

temperature, NDVI),

ASTER (radiometric

surface temperature)

Product SMOS SMOS SMOS Simulated SMOS

Output

resolution (km)

1 10 and 1 4 0.5

Methodology sequential model Build model between

NDVI, LST and soil

moisture

Relationship between

soil moisture and soil

evaporative efficiency

sequential model

Results R2 = 0.70-0.85 in

summer, however

very poor results

obtained in winter

R2 of 0.14-0.21, RMSE

is between 0.9-0.17

Mean slope between

simulated and observed

is 0.94, with an error of

0.012

R2 of 0.68, RMSE of

0.062, bias of 0.045

23

The evaluation of the abovementioned methods highlights significant limitations, which

hinder the successful application of the various downscaling procedures that have been

developed. These include (a) observational days have to be cloud-free, to avoid obscurities in

data retrieval; (b) the presence of vegetation interferes with land surface temperature

retrieval; (c) there is a difference in the input data sensing depth; and (d) the model

assumptions may not be valid in heterogeneous areas (Merlin et al., 2008; 2009; 2010; Piles

et al., 2011).

From the evaluation of these downscaling techniques, several comparisons can be made.

Firstly, all of the downscaling research studies were conducted in the same region; this

simplifies the evaluation process between the different techniques used. It can be noted that

the technique used by Piles et al. (2011) requires the least data input, while the technique

used by Merlin et al. (2009) requires the most input data. All the techniques use SMOS data

or a simulated SMOS data set, as the SMOS satellite is the most recent soil moisture satellite.

The evaluation between the methods is enhanced due to the common aspects of the research

studies.

The differences in the evaluated techniques can be seen in their methodologies, output

resolutions and results. Merlin et al. (2012) had an output resolution of 1 km and showed

good results in summer; however, the technique performed very poorly in winter, when

compared to in-situ soil moisture data. Piles et al. (2011) had an output resolution of 1 and 10

km, but performed poorly, when compared to in-situ soil moisture data. The Merlin et al.

(2010) resulted in the best correlation between observed and simulated soil moisture. Merlin

et al. (2009) showed good correlation and had the finest resolution.

The limitations of the downscaling process can be summarized as follows:

i. The accuracy of the soil moisture product may decrease as the spatial resolution

increases (there is a trade-off between obtaining accurate data and obtaining fine

resolution products);

ii. The process requires input data, which may not be freely or readily available;

iii. The technique may be site-specific;

iv. The complexity of the algorithms used; and

v. Cloud-free images are required, if MODIS products are used.

24

In recent years, the need for fine resolution soil moisture products has resulted in numerous

downscaling research studies being conducted. The main publications in this field have been

studies conducted by Merlin et al. (2009; 2010; 2012) and Piles et al. (2011). In addition to

these, more recent studies have been conducted, which are based on the same principles of

the key techniques. However, there are slight variations, in order to improve and build upon

these key approaches. The new studies include:

i. A study conducted by Zhao and Li (2013) aimed to develop a downscaling method to

improve the spatial resolution of the AMSR-E derived soil moisture product. The

approach was based upon the conventional method of the microwave-optical

synergistic technique, which uses LST, vegetative indexes and albedo. This approach

replaces LST with two temperature temporal variation parameters. The study was

conducted in the Iberian Peninsula for the year 2007. The results showed an

improvement in the approach (R2 increased by 0.08), when the new approach was

compared to the conventional method.

ii. A recent study by Ruiz et al. (2014) was conducted on combining SMOS visible and

near infrared satellite data for high resolution soil moisture over a two-year period in

the REMEDHUS network in Spain. The study used a new downscaling algorithm,

based on a relationship between LST, NDVI and brightness temperature. The study

aimed to obtain a downscaled image with the same accuracy of SMOS, but at the

spatial resolution of MODIS (500 m). The best result obtained was a correlation with

the in-situ measurements of 0.72.

25

4. METHODS OF MODELLING SOIL MOISTURE

Land surface water and surface energy balance models have been used to estimate soil

moisture, as surface soil moisture plays a significant role in controlling the land-atmosphere

water and energy fluxes (Walker and Houser, 2000). Several models have been used to obtain

soil moisture estimates. These include the Variable infiltration capacity (VIC) hydrological

model, the Decision support system for agro-technology transfer crop model (DSSAT), the

climate water budget (CWB), the SEBS, PyTOPKAPI and the Surface Energy Balance

Algorithm for Land (SEBAL) (Meng and Quiring, 2008; Mengistu et al., 2014). The VIC,

DSSAT and CWB are complex models, which require a large amount of input data (Meng

and Quiring, 2008). The SEBAL Model is not freely available to apply due to its intellectual

property rights (Mengistu et al., 2014).

The following section describes the use of the PyTOPKAPI Model soil moisture product and

the SEBS Model to obtain soil moisture estimates. The PyTOPKAPI Model was selected, as

it has been used in South Africa. It is able to provide estimates with a fine spatial and

temporal resolution and the estimates are readily available. The SEBS Model was selected, as

it is an open-source model, it is easy to use and has been validated in several studies

internationally and in South Africa.

4.1 Land Surface Model (PyTOPKAPI)

Soil moisture is a significant parameter in the hydrological cycle and is the prime regulator in

a catchment’s response to precipitation, as it partitions precipitation into infiltration and

surface runoff (Vischel et al., 2008). The topographic kinematic approximation and

integration (TOPKAPI) model was developed by Liu and Todini (2002). The model is fully

distributed and physically-based, as it represents hydrological processes on grid cells, on the

basis of soil physics and fluid mechanics (Vischel et al., 2008). The TOPKAPI Model has

been adapted to perform simulations in land surface modelling mode and its model code has

been adapted, so that it operates as a collection of cells (Sinclair and Pegram, 2010). Each

cell is independent of its neighbours and has a plan area of 1 km x 1 km, with a 0.125o grid

spacing (Sinclair and Pegram, 2010). Each grid cell comprises of three key reservoirs (soil,

overland, channel), as seen in Figure 4.1, which are determined by combing the physically-

26

based mass and continuity equations under the estimate of the kinematic wave model

(Vischel et al., 2008).

Figure 4.1 A schematic of the water transfer in a typical PyTOPKAPI model cell

(Sinclair and Pegram, 2012)

The model is forced by time varying estimates of total evaporation and rainfall for each

model cell (Sinclair and Pegram, 2012). The total evaporation forcing is based on the

modification of the FAO56 reference crop evaporation, whereas the rainfall forcing is applied

in the TRMM 3B42RT product, which is automatically downloaded from the NASA server

and processed locally (Sinclair and Pegram, 2010).

There are six fundamental assumptions on which the TOPKAPI model is based on (Vischel et

al., 2008):

i. Precipitation is spatially and temporally constant over the integration domain;

ii. Unless the soil is already saturated, all precipitation falling on the soil infiltrates;

iii. The slope of the groundwater table corresponds with the slope of the ground;

iv. The saturated hydraulic conductivity in the soil surface layer is constant with depth;

v. Local transmissivity; and

vi. During the transition phase, the variation in water content in time is constant in

space.

27

The TOPKAPI Model has been adapted to South African conditions. A land surface model

was developed, with the final product being the freely available open source software

package, known as the PyTOPKAPI (Sinclair and Pegram, 2013). The PyTOPKAPI Model is

rainfall-runoff model used to examine the soil moisture dynamics at different scales, ranging

from catchment to national scale (Sinclair and Pegram, 2013b).

The PyTOPKAPI Model uses three sets of input data, which consist of the meteorological,

static and remote sensing data sets, as seen in Figure 4.2 (Sinclair and Pegram, 2010). The

meteorological input data include the calculation of reference crop total evaporation and

require parameters, such as relative humidity, temperature, solar radiation flux and wind

speed (Sinclair and Pegram, 2010). The static input data required are the digital elevation

models, land cover and soil properties. The input remote sensing data required are the rainfall

and NDVI products on a three-hour temporal scale. In addition, the solar radiation flux from

the meteorological input data is a satellite-based product and can be considered as a remote

sensing product (Sinclair and Pegram, 2012).

The PyTOPKAPI is an automated modelling system, which produces national estimates of

total evaporation and soil moisture at a three-hour time step over South Africa on a 0.125o

spatial grid (Sinclair and Pegram, 2010).

Figure 4.2 Data-flow diagram showing sources of the dynamic and static data to

produce the main information streams (Sinclair and Pegram, 2010).

28

The purpose of the computation is to obtain a Saturated Soil Index (SSI), which is defined as

the percentage of the soil voids occupied by water.

𝑆𝑆𝐼 = 100 ∗ (Ɵ−Ɵ𝑟

Ɵ𝑠𝑎𝑡−Ɵ𝑟) (4.1)

Where Ɵ is the volumetric soil moisture, Ɵr is the residual soil moisture and Ɵsat is the

saturated soil moisture.

In the original design, Liu and Todini (2002) created the model in such a manner, that

whether the soil store was saturated or not, all rainfall during the computational interval was

added to the soil store. At the end of the interval, an inventory was made and if the store was

full, the excess was assigned to overland and channel stores. Then, based on the content of

the soil store, the drainage was sent to the downslope store (Sinclair and Pegram, 2013a).

This resulted in two effects: the soil store was always depleted at the end of the computation

interval and surface runoff only occurred once the soil store was full, independent of rainfall

intensity (Sinclair and Pegram 2013).

In the adaption of the TOPKAPI Model to South African conditions, the first effect was

cancelled by modifying the continuity equation, so that there was ponding in the overland

store, which resulted in the soil store remaining full after computation (Sinclair and Pegram,

2013a). The second effect was dealt with the introduction of the Green-Ampt infiltration

model, which gave the PyTOPKAPI Model the ability to generate rapid overland flows in

response to intense rainfall events (Sinclair and Pegram 2013).

The benefits of using this model are that the soil moisture estimates obtained have a temporal

resolution of three hours and a spatial resolution grid ≈ 12.5 km. This finer temporal and

spatial resolution better accounts for the heterogeneity of soil moisture, compared to the

current global remote sensing of soil moisture products.

There have been research studies that reviewed the estimation of soil moisture through land

surface modelling.

29

i. Sinclair and Pegram (2010) conducted a study, which aimed to compare soil moisture

estimates from two independent sources over South Africa. The first soil moisture

estimate was provided by automated real-time estimates from the TOPKAPI

hydrological model. The second set of soil moisture estimates was from the ASCAT

remote sensing product.

ii. A similar study was conducted by Vischel et al. (2008) in the Liebenbergslvei

Catchment in South Africa, which compared soil moisture estimates derived from the

TOPKAPI Model with European remote sensing satellite-based soil moisture

estimates. Both studies concluded that the modeled soil moisture measurements

correlated well, when compared to remote sensing soil moisture measurements.

iii. More recent studies, conducted by Sinclair and Pegram (2013), evaluated the

sensitivity of the PyTOPKAPI Model to systematic bias in the variables of soil

properties, evapotranspiration and rainfall. This was achieved by analysing 7200 sites

within South Africa for a two-and-a-half year simulation period. The study concluded

that the model is robust to errors in forcing parameters.

iv. A study conducted by Mengistu et al. (2014) involved the validation of variables of

the PyTOPKAPI Model. The aims of the study were to provide data for the soil

moisture mapping of South Africa, using the PyTOPKAPI Model. The purpose was to

provide accurate field and satellite estimates of total evaporation and soil moisture for

the calibration of the model and to evaluate the spatial variability of soil moisture at a

catchment scale.

4.2 Surface Energy Balance System (SEBS)

Over the past few decades, several methods of estimating evaporation have been developed

and implemented (Ma et al., 2012). The SEBS Model was developed by Su (2002) for the

estimation of atmospheric turbulent fluxes and surface evaporative fraction, using remote

sensing data (Su et al., 2003). The SEBS Model is physically-based, has the advantage of

being less site-specific, open-source, it has been applied extensively (especially as an

academic tool for research purposes) and does not require subjective intervention by the

model user (Gokman et al., 2012).

30

The model requires three sets of informational input data. The first set comprises of albedo,

temperature, leaf area index and vegetation coverage (Su, 2002). In cases where this

information is unavailable, the NDVI and the vegetation height are used instead. These

informational inputs can be obtained from remote sensing data, in conjunction with additional

land surface information. The second set of information comprises the reference boundary

layer height (Su, 2002). The third set of input data required is the downward solar and long-

wave radiation, which are obtained either from direct measurement, parameterization or

model output (Su, 2002).

The SEBS Model consists of numerous individual components, to estimate the net radiation

and soil heat flux and to partition the available energy into latent and sensible heat fluxes.

The Surface Energy Balance is expressed as follows (Equation 4.2) (Wang and Li, 2011):

Rn = G𝑜 + H+ λE (4.2)

Where Rn = net radiation, Go = soil heat flux, H = sensible heat flux and λE = latent heat flux.

The considerations of the surface energy balance, at limiting cases, are used to determine the

relative evaporation (Wang and Li, 2011). When the dry limit is considered, the latent heat is

zero and the sensible heat flux is at its maximum (Wang and Li, 2011):

λEdry = Rn − Go − Hdry ≡ 0

Hdry = Rn − Go (4.3)

When the wet limit is considered, evaporation occurs at a maximum rate, whilst the sensible

heat flux takes the minimum value (Wang and Li, 2011):

λEwet = Rn − Go − Hwet (4.4)

The relative evaporation can be expressed as (Su et al., 2003):

31

Λr = λE

λEwet= 1 −

λEwet − λE

λEwet (4.5)

By substituting Equations 4.2 – 4.4 in Equation 4.4 and rearranging (Su et al., 2003):

Λr = 1 − H− Hwet

Hdry− Hwet (4.6)

Applying the mass conservation principle and integrating with respect to depth and time

increments (Su et al., 2003):

∫ θ(z, t2)dz − ∫ θ(z, t1)dz = Q(z1) − Q(z2)z2

z1

z2

z1 (4.7)

where θ is the volumetric soil moisture content, t is the time and z is the vertical distance.

Applying Equation 4.7 with boundary conditions Q (z1) = Po + Io –E at the soil surface and Q

(z2) = Ic at the bottom of the rooting zone. The change in soil water content can be expressed

as (Su et al., 2003):

Θ(t2) − Θ(t1) = Po + Io + Ic − E (4.8)

where Θ is the volumetric soil water content in the rooting zone, Po is the precipitation, Io is

the irrigation, Ic capillary flux and E is evaporation. The water balance is then considered at

limiting cases. The wet limit is saturation, so that Θ (t1) = Θwet. At the dry limit Θ (t2) = Θdry,

the evaporation is zero. From Equation 4.8 (Su et al., 2003):

Θwet − Θdry = Icwet − Ewet (4.9)

For any time between the two boundary conditions:

Θ − Θdry = Icwet − E (4.10)

Manipulating Equation 4.9 and Equation 4.10 (Su et al., 2003):

Θ − Θdry

Θwet− Θdry=

E−Ic

Ewet−Icwet (4.11)

32

It is assumed that the capillary flux is linked to the soil texture and is less than that of the

uptake of root water (Ic = Iwet). By defining Rθ = Θ/ Θwet as the relative soil water content and

using Equation 4.11 (Su et al., 2003):

Rθ = Θ

Θwet=

λE

λEwet (4.12)

This inverse relationship is often used in hydrological modelling and can be expressed as:

λE

λEwet= f (

SMactual

SMfield capacity) = Ʌ (4.13)

The relative evaporation is therefore calculated as the relative soil moisture contained in the

rooting zone. The term f is used as a linear relationship. The SM field capacity can be inferred

from the soil texture.

An alternative method, which uses an empirical relationship between the evaporative fraction

and soil moisture, was developed by Bastiaanssen et al. (1997). The equation is based on two

large-scale climate-hydrology studies, which investigated the interactions between soil

moisture, evaporation and biomass (Scott et al., 2003; Bezerra et al., 2013). The empirical

equation was modified by Scott et al. (2003), by standardizing soil moisture with saturated

soil moisture, to create relative soil moisture content (Scott et al., 2003):

θ

θsat= exp[(ᴧ−1)/0.421] (4.14)

These methods of soil moisture estimation result in a finer spatial resolution than that of

current remote sensing soil moisture products, as the remotely sensed data is observed at a

finer spatial scale. In addition, the soil moisture being estimated is at a greater depth than that

estimated by microwave remote sensing (Bezerra et al., 2013). The limitation of this method

is that the observations and measurements are restricted to cloudless days, as retrieval is

affected by atmospheric conditions. Therefore, only days which are cloud-free can be used, in

order to reduce measurement errors.

33

The SEBS Model has been applied globally in several studies (Wang and Li, 2011; Gokman

et al., 2012; Ma et al., 2012).

i. A study conducted by Su et al. (2003) derived a measure for the quantification of

drought severity. The study used the SEBS Model, meteorological data and NOAA

satellite images to derive relative evaporation in North China. A drought severity

index (DSI) was subsequently derived from the relative evaporation. The DSI gives

an indication of the soil moisture status, so that a high relative evaporation is

associated with a high soil moisture status. The study concluded that the DSI and

actual soil moisture measurements show a good comparison.

ii. Studies conducted by Scott et al. (2003) and Bezerra et al. (2013) evaluated the

relationship to calculate soil moisture from remote sensing data. The study conducted

by Bezerra et al. (2013) was carried out in the Apodi Plateau in Brazil. The study

concluded that there was a close correlation between measured and estimated soil

moisture value (R2=0.84).

34

5. SYNTHESIS OF LITERATURE

The literature review detailed in-situ, remote sensing and modelled soil moisture

measurement and estimation methods. Soil moisture is an important parameter, which has

several implications for a number of applications. Conventional in-situ methods have been

invaluable in providing validation and calibration data. The major limitation, however, is that

the point measurements do not represent the spatial characteristics of soil moisture.

In recent years, the CRP technique has been used to provide continuous measurements of

area-averaged soil moisture at an intermediate scale, which bridges the gap between point

measurements and large footprints obtained from remote sensing.

Remote sensing of soil moisture is a promising technique, as it accounts for the spatial and

temporal characteristics of soil moisture. It has numerous advantages; however, its

application is limited, due to its coarse resolution. The product can be downscaled, in order to

obtain a spatial resolution applicable for use in hydrological applications. There are several

downscaling techniques that vary in complexity and the choice of the downscaling method

used will depend on the intended output resolution, the input data available and the general

ability of the researcher. Although a finer spatial resolution is favourable, the loss of remote

sensing soil moisture product accuracy, due to downscaling, is a concern.

Soil moisture is an important parameter in both the land surface water and land energy

balance and can be estimated from modelling. The PyTOPKAPI rainfall-runoff model has

been used with confidence to estimate soil moisture across South Africa.

The SEBS Model can also be used to estimate soil moisture. Both models use remote sensing

data as input data. The advantage of using models to estimate soil moisture is that the spatial

resolution is greater than that of current global remote sensing products.

There needs to be a shift from using point measurements, to using area-averaged soil

moisture measurements, to calibrate and validate remote sensing and model products. This

will aid in reducing the vertical and horizontal scaling issues. There have been no studies

35

conducted in South Africa on the comparison of soil moisture estimate methods, using the

CRP as validation data for satellite-based soil moisture estimates.

From the literature reviewed, the gaps in past research studies have become apparent. A

major gap has been the use of point measurements for the calibration and validation of

remote sensing and modelled soil moisture estimates. This has resulted in horizontal and

vertical scaling issues, therefore there is a need for a shift towards the use of area-average

calibration and validation data, in order to bridge the gap between scales.

Downscaling procedures have not been used in South African case studies; however, they are

essential in obtaining remote sensing soil moisture estimates at a fine spatial resolution,

which can be used in hydrological applications. Where downscaling has been used, the

downscaled estimates were compared to point measurements.

The PyTOPKAPI Model has been implemented in South Africa with confidence; however,

its estimates have not been evaluated against area-averaged soil moisture measurements.

There have been no South African case studies on the CRP, therefore a comprehensive

evaluation of the current methods of soil moisture against the area-averaged soil moisture

measurements obtained from the CRP, need to be conducted in South Africa.

36

6. METHODOLOGY

Several studies referred in the literature review component (chapters two, three and four)

have suggested that the current methods of soil moisture estimation do not adequately

provide measurements for critical hydrological applications, as each method has its

limitations. Although the current methods can provide accurate estimates of soil moisture, the

representativeness of these methods may vary at different spatial scales. These methods need

to be evaluated and improved, in order to obtain soil moisture data that can be used in critical

hydrological applications. Further research is warranted. The following research questions

have been formulated to address both the general and specific objectives of the study:

How suitable is the CRP technique in providing spatial estimates of soil moisture?

How accurate are the Level Three AMSR2 and SMOS soil moisture estimates?

How suitable are the PyTOPKAPI Model product and the back-calculation using the

SEBS model in providing estimates of soil moisture?

What are the effects of seasonality on the different methods of soil moisture

estimation?

The general objectives of this study were to calibrate and validate the CRP and subsequently

use the calibrated CRP soil moisture estimates to validate satellite-based soil moisture

estimates. This involved setting up the CRP in Cathedral Peak Catchment VI. The CRP

continuously monitored soil moisture at an hourly time-step. During this time, four

calibrations were undertaken to determine the representative gravimetric water content of the

study area on that calibration day. An in-situ soil moisture network was also set up that

consisted of TDR and Echo probes. The aim of the in-situ soil moisture measurements was to

provide data that could be used to create a representative soil moisture dataset of the study

area, which could then be used to validate the CRP.

Calibrated CRP estimates were then used to validate satellite-based soil moisture estimates.

These consisted of remote sensing products, back-calculated soil moisture using outputs from

the SEBS model and modelled soil moisture product from the PyTOPKAPI model. Although

the downscaling of remote sensing data was reviewed and evaluated in the literature, the

procedure of downscaling remote sensing products did not fall within the scope of this study

and therefore, was not applied.

37

6.1 Study Site

The study site for this research was Cathedral Peak Research Catchment VI. Cathedral Peak

was established in 1935, as the chief centre for forest hydrological research in the summer

rainfall region (Scott et al., 2000). Its primary purpose was to examine the influences of

various management practices on the vegetation and yield of the local mountainous

catchments (Everson et al., 1998). Cathedral Peak lies within KwaZulu-Natal, in the Tugela

Catchment, as shown in Figure 6.1. It comprises of fifteen gauged catchments that are

situated on the little Berg, located below the Drakensberg escarpment, which creates a natural

border between the north-eastern side of Lesotho and South Africa (Gush et al., 2002).

Figure 6.1 Location of Cathedral Peak Catchment VI, within the Tugela catchment,

within KwaZulu-Natal

The Drakensberg mountain range is the highest mountain range in South Africa and gives rise

to many of the rivers, which are of great economic importance to the country (Everson et al.,

1998). The fifteen research catchments (numbered I to XV) received different land

38

treatments, such as varying burning regimes, grazing, afforestation and protection from

burning (Kuenene et al., 2009).

Catchment VI is the focal catchment in this study. It has a catchment area of 0.68 km2 and is

located by latitude of 28.99o S and longitude of 29.25o E. Catchment VI is moderately

dissected by streams and has a stream density of 3.25 km/km2 (Everson et al., 1998). The

altitude ranges from 1860 m at the weir (northern-most point of the catchment) to 2070 m at

the highest point of the catchment (Kuenene et al., 2009). The average slope of the catchment

is 19%. A topographic map of Catchment VI can be seen in Figure 6.2.

Figure 6.2 Topographic map of Catchment VI

The landcover of Catchment VI is Highland Sourveld grassland (Everson et al., 1998). The

soils in the catchments are moderately weathered immature soils, which are primarily derived

from basalt (Scott et al., 2000). The soils in the catchment are classified as lateritic yellow

and red earths, with heavy black soils occurring in saturated zones and along stream banks

(Jarmain et al., 2004). There is a contrast in soil properties among the soil layers. The topsoil

has a friable consistency, which results in rapid infiltration, whilst the subsoil has a very high

clay content, which results in poor infiltration. The topsoil has a high organic matter content

(6-10%), which results in a high water-holding capacity (Kuenene et al., 2009). The region is

39

characterized climatically by its cold dry winters and hot wet summers. The mean annual

precipitation is 1400 mm, with 85% falling between October to March (Gush et al., 2002).

Catchment VI has a mean annual precipitation of 1299 mm (Jarmain et al., 2004).

Catchment VI is one of the smallest research catchments in Cathedral Peak. It is 4.027 km

south-east of the Mikes Pass weather station and is the most instrumented catchment. It

contains the CRP, eddy covariance system, large aperture scintillometer, weather station, a

compound V-notch weir and a tipping bucket rain gauge.

6.2 Cosmic Ray Probe

The CRP is a relatively new and innovative in-situ instrument used to obtain area-averaged

soil moisture measurements.

Set up of the cosmic ray probe

The CRP was installed on the 28th of February 2014 in Cathedral Peak research catchment VI

(28.99o S and 29.25o E), as shown in Figure 6.3. The CRP was positioned towards the centre

of the catchment and due to its measurement footprint, the soil moisture measurements

obtained are assumed to be representative of the catchment area.

Figure 6.3 CRP in Cathedral Peak Catchment VI

40

The CRP method requires calibration to obtain accurate soil moisture estimates. The

calibration process involves obtaining an estimate of the area-averaged soil moisture content

over the CRP measurement footprint by gravimetric sampling. Soil samples were taken at

three radial rings, extending outwards from the CRP. The radial rings were situated at a

distance of 25 m (A), 100 m (B) and 200 m (C) from the CRP, as shown in Figure 6.4. At

each of the three rings, eight points were taken at an equal distance along the circumference

of the ring. Therefore, points were taken at 0, 45, 90, 135, 180, 225, 270 and 315 degrees

from the CRP.

The location of these 24 points (3 rings x 8 points) were mapped, using GIS software and

their coordinates were recorded (Table 6.1). These points are located within Catchment VI, as

illustrated in Figure 6.5.

A1

A2

A3

A4

A5

A6

A7

A8

B1

B2

B3

B4

B5

B6

B7

B8

C1

C2

C3

C4

C5

C6

C7

C8

CRP

25m

100m

200m

CRP

Figure 6.4 Diagram of sampling points

41

Table 6.1 Geographical coordinates of sample points

Sample Latitude Longitude Sample Latitude Longitude Sample Latitude Longitude A1 -28.9929 29.2519 B1 -28.9922 29.2519 C1 -28.9913 29.2519 A2 -28.9929 29.2520 B2 -28.9924 29.2526 C2 -28.9918 29.2533 A3 -28.9931 29.2521 B3 -28.9931 29.2529 C3 -28.9931 29.2539 A4 -28.9932 29.2520 B4 -28.9937 29.2526 C4 -28.9944 29.2533 A5 -28.9933 29.2519 B5 -28.9940 29.2519 C5 -28.9949 29.2519 A6 -28.9932 29.2517 B6 -28.9937 29.2511 C6 -28.9944 29.2504 A7 -28.9931 29.2516 B7 -28.9931 29.2508 C7 -28.9931 29.2498 A8 -28.9929 29.2517 B8 -28.9924 29.2511 C8 -28.9918 29.2504

Figure 6.5 Position of calibration sample points within Cathedral Peak Catchment VI

C1

B1

A1

A5

B5

C5

A3 B3 C3 A7 B7 C7

42

Field sampling

Each sample point was entered and saved in a hand-held GPS system, in order to pinpoint the

location of the sample points that had been determined. Once each point was found, they

were visually marked, by inserting a metal rod into the ground and tying a piece of hazard

tape to the top of the rod. At each point, soil samples were taken and labelled X, Y and Z. At

each of these three points, an auger and garden trowel were used to obtain soil samples at

depths of 0.05, 0.10, 0.20 and 0.30 m and these were placed in separate plastic bags. The

plastic bags were clearly labelled according to the ring (A, B, C), sample point in the ring (1,

2, 3, 4, 5, 6, 7, 8), replicate (X, Y, Z) and depth (5, 10, 20, 30), as seen in Figure 6.6. The soil

samples obtained were properly sealed, stored in a box and transported to the laboratory.

Figure 6.6 Field samples contained in plastic bags

Gravimetric water content determination

The determination of the gravimetric soil moisture content of each sample point was required

to obtain a representative gravimetric soil moisture content of the study area, for the

calibration of the CRP. The soil sample plastic bags were re-opened and a 20 g sub-sample of

each sample was weighed, using a mass balance, its weight recorded to three decimal places

and placed onto pre-weighed individual foil trays, which were labelled according to the

sample (Figure 6.8). The samples were then placed in an oven at 105oC for 24 to 48 hours,

after which, the samples were removed from the oven, weighed and recorded once more. The

43

gravimetric water content was then calculated on a wet basis, using the gravimetric water

content:

Gravimetric water content = Wet mass−Dry mass

Dry mass × 100 (6.1)

Figure 6.7 Weighing the soil samples and placing them in the oven

Four calibrations were carried out over a period of eight months. The calibration dates were

the 9th of July 2014, the 28th of August 2014, the 2nd of December 2014, and on the 22nd of

January 2015.

Creating a soil moisture network

It was essential to create a soil moisture network comprising of in-situ soil moisture

measurement instruments at several locations within the CRP measurement area. This was

required in order to obtain data that could be used to validate the CRP measurements. There

were three sets of measurement instruments installed between the 9th and 10th of July 2015. A

soil pit was dug and TDR probes were inserted horizontally at depths of 0.05, 0.10, 0.15, 0.20

and 0.30 m, as illustrated in Figure 6.7 (a). Five wireless TDR sensors (0.12 m) were inserted

vertically into the soil surface at sample points A5, B5, C5, A7 and B7, as illustrated in

Figure 6.7 (b). Shallow soil pits were dug at A1, B1, C1, A3, B3, C3, B7 and C7, and Echo

probes were inserted horizontally at a depth of 0.10 m (Figure 6.7 (c)). The echo probes were

buried and their data loggers were attached to the metal rod at the subsequent sample points

(Figure 6.7 (d)).

44

(a) (b)

(c) (d)

Figure 6.8 (a) Setting up the TDR pit, (b) Wireless TDR, (c) Echo probe and (d) Data

Hobo Onset logger for the Echo probe

The wireless TDR sensors, which were placed on the 10th of July 2014 at points A7, A5, B5,

B7 and C5, were removed on the 28th of August 2014 to check a problem with the data

logging program.

Catchment burning

Catchment VI is subjected to prescribed burning for fire protection and management

purposes, as it reduces the amount of groundcover vegetation during the drier (cooler) months

and decreases the risk of wildfires. These wildfires could cause greater damage later on in the

season. Without fire, grasslands could transform into shrub-lands or forests. Prescribed

burning is different to wildfires and has less of an environmental impact. Prescribed burning

is undertaken by a team of skilled and equipped laborers, who use firebreaks to control the

burning stretch. There are three main types of burning strategies used:

45

i. Back-burning is the technique of setting small fires along a fire-break and burning

back to the main fire front. This burning type is usually against the ground-level

winds and often downslope, which results in a lower intensity fire that moves at a

slower speed, which makes the fire easier to control.

ii. Head-burning is the technique of burning with the ground-level wind direction and is

usually upslope. This burning type has a high intensity and moves at a faster rate,

which makes it harder to control.

iii. Flanking is the technique of setting a fire, which moves parallel to and into the wind.

This burning type is used to supplement the other two burning techniques.

The hydrological consequences of fire include the removal of vegetation cover and the direct

effect of the fire heating the soil (Scott, 1994). The removal of the vegetation cover could

have an impact on the soil moisture status, as the soil surface has greater exposure to

atmospheric conditions. This could lead to increased soil evaporation, which could result in

drier topsoils (Stoof et al., 2012). Conversely, there could be an increase in soil moisture in

the subsoils, due to a reduction in transpiration (Stoof et al., 2012). The removal of

vegetation may also result in increased erosion and sedimentation, due to an increased

exposure of the soil surface to raindrop impact (Scott, 1994). There is a possibility of post-

fire pore-clogging by infiltrated ash, which leads to reduced infiltration rates and causes an

increase in overland-flow (Stoof et al., 2012). There could also be the development of soil

water repellency during and after the fire.

Although these hydrological consequences have been reported, the effects of fire on

hydrological responses at a catchment scale are minor, as these fires are generally prescribed

burns and the issue of scale plays a significant role in the overall change detection of the

hydrological system (Scott, 1994).

On the 5th of September, a scheduled burning from an adjacent Catchment (VII) spread and

caused the unscheduled burning of Catchment VI. As a result of the fire, the Echo probe soil

moisture sensors were destroyed. However, the data from the hobo loggers were retrieved

(Figure 6.9). The larger equipment in Catchment VI were not harmed due to the fire breaks,

as seen in Figure 6.10. Due to the loss of the echo probe sensors, the wireless TDR sensors

had to be placed, to best represent the catchment. This was done by identifying the wet,

46

intermediate and dry areas of the catchment, by using the soil moisture map obtained from

the first calibration.

Figure 6.9 Data retrieved from burned echo probe data loggers

Figure 6.10 Protection of equipment by fire breaks in Catchment VI

Bulk density determination

The gravimetric determination technique was used for the calibration of the CRP. The

gravimetric soil moisture content is expressed, by weight, as the ratio of the mass of water

present to the dry weight of the soil sample (g/g). The CRP, however, measures the

volumetric soil moisture content, which is expressed, as the ratio of volume of water to the

total volume of the soil sample (cm3/cm3). This is further expressed as volumetric water

content in percentage.

47

The calibration of the CRP requires a representative bulk density value, to convert the

gravimetric soil moisture content to volumetric water content. In order to convert gravimetric

soil moisture into volumetric soil moisture, the gravimetric water content is multiplied by the

bulk density. The bulk density determination was carried out on the 2nd of December 2014.

Undisturbed soil cores were taken at three locations in the catchment, corresponding with

three sample points (C1, CRP and C5) and at three depths (0.00 – 0.05 m, 0.05 – 0.10 m and

0.10 – 0.15 m). These sites were selected to best represent the catchment.

At each of the three sites, steel cylindrical cores were inserted vertically into the soil. Starting

with the 0.00 – 0.05 m measurement, the steel core was placed on the surface of the soil. A

wooden board was placed on the top of the core and the core was hammered down into the

soil. The wooden board was used to distribute the force of the hammer on the steel core, so

that the core moves down equally. Once the steel core had reached its intended depth, the soil

around the core was loosened and removed carefully. The steel core was then removed,

clearly labeled and its ends tightly sealed. The same procedure was then carried out at 0.05 –

0.10 m and 0.10 – 0.15 m depths.

The soil cores were then transported to the laboratory, where they were opened carefully. A

knife was used to trim the excess soil of the ends of each side of the core, so that the volume

of the soil equaled the volume of the core. The soil cores (steel core and soil) were oven-dried

at 105oC for 48 hours. The mass of the soil cores were weighed and recorded after oven

drying (refer to Figure 7.14). The mass of the cylindrical steel cores were recorded and

subtracted from the soil core mass, to obtain the mass of dry soil. The volumes of the soil

cores were determined by the dimensions of the cylindrical steel cores. The volume of the

steel core is the volume of the soil.

The bulk densities were calculated, using Equation 6.2:

Bulk Density (g/cm3) = 𝐷𝑟𝑦 𝑀𝑎𝑠𝑠 𝑜𝑓 𝑠𝑜𝑖𝑙 (𝑔)

𝑇𝑜𝑡𝑎𝑙 𝑉𝑜𝑙𝑢𝑚𝑒 (𝑐𝑚3) (6.2)

The calculated bulk densities of each sample point and depth along with the variables used to

calculate them are illustrated in Table 6.2.

48

Table 6.2 Variables required in obtaining the bulk density

Sample Depth (m) Ms (g) Volume (cm3) Bulk Density (g/cm3)

C1 0.05 38.863 98.175 0.396

C1 0.10 50.404 98.175 0.513

C1 0.15 56.127 98.175 0.572

CRP 0.05 55.45 98.175 0.565

CRP 0.10 60.359 98.175 0.615

CRP 0.15 65.143 98.175 0.664

C5 0.05 60.253 98.175 0.614

C5 0.10 65.194 98.175 0.664

C5 0.15 71.86 98.175 0.732

It can be seen that, as the depth increases, the bulk density increases. The calculated bulk

density values were low. This can be attributed to the soil cover, organic matter content, soil

structure, porosity and the lack of compaction, as the catchment is situated in an undisturbed

area. Initially, the calibration involved a depth weighting soil moisture procedure. Therefore,

the average bulk density per depth was required. The new and current calibration procedure

does not involve a depth weighted soil moisture, thus one representative bulk density value

was required and was calculated by averaging all the bulk density values. The average

measured bulk density was thus 0.593 g/cm3.

6.3 Calibration

In total, four calibrations were completed. Two of the calibrations were completed in the

“dry” period and two were completed in the “wet” period. Subsequently, one “full”

calibration and one “partial” calibration were completed in each period. The partial

calibrations were the result of limited field time. The calibrations, number of samples and

depths obtained, are shown in Table 6.3. The full calibrations (1 and 3) have a large number

of samples, due to all 24 points being covered. Furthermore, three replicate samples were

obtained at each point, at each depth. The first calibration involved collecting samples from a

depth of 0.30 m; however, the CRP did not measure the soil moisture at a depth greater than

0.20 m, due to the relatively high soil moisture in the “dry” period. Therefore, this depth was

discontinued for the subsequent calibrations.

49

Table 6.3 Information regarding the calibration sampling

Depth Measured (m)

Calibration Date Period No. of samples 0.05 0.10 0.20 0.30

1 (Full) July 9, 2014 Dry 288 x x x x

2 (Partial) August 28, 2014 Dry 36 x x x -

3 (Full) December 2, 2014 Wet 216 x x x -

4 (Partial) January 22, 2015 Wet 36 x x x -

For the purpose of the calibration of the CRP, an average (one single) soil moisture value per

calibration is needed, thus a representative soil moisture value was determined for each

calibration.

Soil moisture maps of each calibration were created by averaging the depths of each sample

point and using the kriging interpolation technique to create soil moisture maps between the

sample points obtained. These maps were not required for the calibration procedure, but were

created to show the spatial characteristics of soil moisture in the research area. The soil

moisture map values are in percentage volumetric content.

Calibrations one and three are more detailed, as they have more points. Calibrations one and

two were done in the dry period (Figure 6.11 and Figure 6.12). Calibrations three and four

where done in the wet period (Figure 6.13 and Figure 6.14). From these soil moisture maps,

general trends can be seen. The soil moisture is lowest in the south (upslope) and the highest

in the north (downslope). It can be seen that topography affects the soil moisture content, as

the soil moisture content generally increases as the altitude decreases along the north-south

transect. The change in soil moisture in the research area, illustrates the limitation of using in-

situ point measurements due to the spatial variability of soil moisture.

50

Figure 6.11 Soil moisture map of the first calibration (9th of July 2014)

Figure 6.12 Soil moisture map of the second calibration (28th of August 2014)

51

Figure 6.13 Soil moisture map of the third calibration (2nd of December 2014)

Figure 6.14 Soil moisture map of the fourth calibration (22nd of January 2015)

52

Soil moisture varies both spatially and temporally. With regards to its spatial variability, soil

moisture varies horizontally, as well as vertically. The horizontal and vertical spatial

variability can be due to several factors, such as topography, rainfall distribution, soil

characteristics and vegetation. If calibration one is considered, for each site (A1, A2, A3,

etc.), three auger points were taken, so that there are three replicate samples at each sample

point. The data of sample point A1, from calibration one, is depicted in Figure 6.15. From

Figure 6.15, it can be seen that soil moisture varies with depth. In general, the soil moisture

increases with depth. It is also evident that, although the three auger points were not more

than one meter apart from one another, there was a noticeable change in the soil moisture

content.

Figure 6.15 Volumetric water content against depth for each replicate at one sample

point

The averages of all 24 soil sample points, for calibration one, are plotted against depth on the

same graph. In Figure 6.16, it is evident that the change in soil moisture content is not

constant. In general, the soil moisture seemed to increase with depth. The 24 plotted points

show the variability of soil moisture of the area. The graph also shows the variability of soil

moisture with depth, as some curves show a noticeable change in soil moisture with depth,

whilst others have a fairly constant soil moisture content down a profile.

5

10

15

20 25 30

De

pth

(cm

)

VWC (%)

Replicate1

Replicate2

Replicate3

Average

53

Figure 6.16 Volumetric water content against depth for all 24 sample points

The calibration procedure was carried out once all four calibrations had been completed. The

aim of the calibration was to determine the average No value, by averaging the No value

determined for each calibration. The data from the CRP is sent, via satellite link, to the

COSMOS server: http://cosmos.hwr.arizona.edu/Probes/probemap.php. The data is available

at three levels. Level one data comprises of raw counts of fast and thermal neutrons. Level

two data is quality controlled, uses level one data and subsequently converts it into a suitable

format to determine soil moisture. Level three data takes level two data and converts it

directly to measurements of soil moisture. The calibration procedure outline that was

followed is that of Franz (2014) and Franz et al. (2015).

The first step in the calibration procedure was to correct the neutron counts. This involved

determining the neutron correction factors, using the following equations (Franz, 2014):

N = N′∗ CP ∗ CWV

CI ∗ CS (6.3)

where N is the corrected neutron counts per hour, N’ is the raw moderated neutron counts, CP

is the pressure correction factor, CWV is the water vapour correction factor, CI is the high-

energy intensity correction factor, and CS is the scaling factor for geomagnetic latitude

(Franz, 2014).

5

10

15

15 25 35 45 55

De

pth

(cm

)

VWC (%)

54

CP = exp (Pi− Po

130) (6.4)

CI = N H

i (t)

N H o (6.5)

CS = f(x, y, z, t) (6.6)

Where x, y, z, is location and elevation, and t is time.

CWV = 1 + 0.0054 (ρvi (T, P, RH) − ρv

o(T, P, RH)) (6.7)

where rvi is the absolute humidity of the air (g/m3), rv

0 is the reference absolute humidity of

the air (g/m3), T is air temperature in (oC), P is pressure (mb), and RH is relative humidity

(%) (Franz, 2014).

For the calibration procedure, the level two data was used, which has already been corrected

for CP, CI and CS. If level one data were used, then the original equation and all the

corrections are required. The revised neutron count correction equation:

𝑁 = 𝑁′ ∗ 𝐶𝑊𝑉 (6.8)

Therefore, when using level two neutron counts, only the correction for absolute water

vapour (CWV) is required. The absolute water vapour calculation was obtained, using the

following equations (Franz, 2014):

eso = 611.2 ∗ exp (17.67∗T

243.5+T) (6.9)

Where eso is the saturated vapour pressure at surface (Pa) and T is air temperature (oC)

(Franz, 2014).

eo = RH

100∗ es𝑜 (6.10)

55

Where eo is actual vapour pressure at surface (Pa) and RH is the relative humidity (%) (Franz,

2014).

ρv = eo

Rvap∗(T+273.15)∗ 1000 (6.11)

Where rv is the absolute humidity of air (g/m3), Rvap =R

0.001Mvap

is the gas constant for

water vapour (J/K/kg), R is universal gas constant (= 8.31432 J/mol/K), Mvap is the molar

mass of water vapour (= 18.01528 g/mol = 0.01801528 kg/mol ), and T is air temperature

(oC).

These equations require hourly temperature and relative humidity data for the same time

period as the level two neutron count data (hourly). The data set required covered the period

of March 2014 to March 2015. Although the CRP system does measure temperature and

pressure, the sensors are inside the data logger casing, which results in the measurement of

the loggers conditions. The primary data set used was from the eddy covariance system;

however, this only had data from the 12th of July 2014 and there were a few gaps in the data

set, which are illustrated by the vertical red arrows, as seen in Figure 6.17. This data alone

could not be used in the calibration, as the time period that it covers is not adequate and the

gaps are problematic.

Figure 6.17 Hourly temperature (The same data gaps present in the relative humidity

dataset)

0

5

10

15

20

25

30

35

3/1/2014 4/30/2014 6/29/2014 8/28/2014 10/27/2014 12/26/2014 2/24/2015

Tem

pe

ratu

re (

oC

)

Date

56

In order to calibrate the CRP, a complete and reliable temperature and relative humidity data

set was required. Therefore, the data from the eddy covariance system was in-filled and

patched with data from the nearby Mikes Pass weather station (refer to Figure 6.18). The

Mikes Pass weather station had the hourly temperature and relative humidity for the period

required.

Figure 6.18 Complete daily air temperature data for Cathedral Peak Catchment VI

Once the hourly CWV values were determined, the corrected neutron count rates were

determined. The following calibration function was then used to determine the No value for

each calibration (Franz et al., 2015):

(θP + θLW + θSOC)ρbd = 0.0808

N

No−0.372

− 0.115 (6.12)

Where q p is pore water content (g/g), qLW is lattice water content (g/g), qSOCeq is soil organic

carbon water content (g/g), rbd is dry soil bulk density (g/cm3), N is the corrected neutron

counts per hour, and N0 is an instrument specific calibrated parameter that represents the

count rate over dry silica soils.

θSOC = (TC − 12

44 CO2) 0.5556 (6.13)

0

5

10

15

20

25

30

35

3/1/2014 5/30/2014 8/28/2014 11/26/2014 2/24/2015

Tem

pe

ratu

re (

oC

)

Date

57

Where TC is the soil total carbon (g/g), CO2 is the soil CO2 (g/g), 12/44 represents the

stoichiometric ratio of carbon to CO2, and 0.5556 is the stoichiometric ratio of H2O to

organic carbon (assuming organic carbon is cellulose C6H10O5).

θsoc was not determined, but was given a value of 0.01 g/g.

θlw was determined to be 0.15433 g/g. A 50 gram representative soil sample was sent to

Activation Laboratories in Canada for θlw determination.

There is a need to correct for biomass (Franz et al., 2015)

(θP + θLW + θSOC)ρbd = 0.0808

N

No(BWE)−0.372

− 0.115 (6.14)

where BWE is the biomass water equivalent (mm). The biomass calculation is done for

vegetation types, whose biomass changes with their growing stage. Due to the vegetation of

Catchment VI being grassland, the biomass did not change much and the change in biomass

was therefore insignificant and a biomass correction was not required. The burning of the

catchment affected the biomass content greatly. This issue was resolved by performing a

calibration before (calibration two) and a calibration after (calibration three) the burning.

The neutron count (N) for each calibration was determined to be the average neutron count,

during which the soil samples for that calibration were obtained (Table 6.4). For the “full”

calibration, the duration was ≈ six hours. The duration of the “partial” calibration was ≈ three

hours

Table 6.4 The gravimetric soil moisture, bulk density and Neutron count

Calibration

Gravimetric Soil Moisture

(g/cm3)

Bulk density

(g/cm3)

Neutron Count

(count/hr)

1 0.490 0.593 1731.684

2 0.438 0.593 1761.408

3 0.647 0.593 1652.600

4 0.741 0.593 1611.059

58

The CRP data prior to calibration (blue), with the calibration points (red), were plotted

against time:

Figure 6.19 CRP soil moisture estimates prior to calibration, with calibration points

From Figure 6.19, it can be seen that the soil moisture calibration values correlate well with

the soil moisture estimates of the CRP. This is the result of the uncalibrated No value being

relatively close to the actual No value. The uncalibrated No value (3000) was set by the

developers of the probe.

The calibration equation (Equation 6.12) was used to determine the No value for each

calibration, which are shown in Table 6.5.

Table 6.5 Date and calculated No value for each calibration

Calibration Date No

1 09 July 2014 3250.573

2 28 August 2014 3242.507

3 02 December 2014 3255.973

4 22 January 2015 3248.243

The average No value for the calibrations was calculated to be 3249.324. This calculated No

value was then used in the calibration function equation (Equation 6.12), to determine the

volumetric soil moisture content.

0

10

20

30

40

50

60

70

80

90

3/1/2014 4/30/2014 6/29/2014 8/28/2014 10/27/2014 12/26/2014 2/24/2015

VW

C (

%)

Date

59

The calibration function was used to determine an hourly (θp+θlw+θsoc)ρbd, by using a constant

No and the hourly corrected N values. The lattice water and soil organic carbon water

(θlw+θsoc) were subtracted from (θp+θlw+θsoc)ρbd to determine the Volumetric Water Content

(VWC) (calibrated CRP dataset). The hourly calibrated VWC from the CRP was plotted

against time, as shown in Figure 6.20.

Figure 6.20 Hourly soil moisture estimates of the calibrated CRP

From Figure 6.20, it can be seen that the calibrated hourly CRP follows the same trend as the

uncalibrated hourly CRP dataset (shown in Figure 6.24); however, there is a downward shift

in the points. The neutron count was then plotted against the calibrated VWC. The

exponential function from the graph (Figure 6.21) can be used to determine the VWC from

the neutron count measured.

Figure 6.21 Neutron count against volumetric water content

0

10

20

30

40

50

60

70

80

90

3/1/2014 4/30/2014 6/29/2014 8/28/2014 10/27/2014 12/26/2014 2/24/2015

VW

C (

%)

Date

0

10

20

30

40

50

60

70

80

90

1400 1500 1600 1700 1800 1900 2000

VW

C (

%)

Neutron (count/hr)

60

The hourly CRP data was converted into a daily dataset (Figure 6.22). From Figure 6.22, it

can be seen that the conversion from hourly to daily data results in a smoother dataset, as the

fluctuations are averaged. The soil moisture dataset generally shows that soil moisture is

higher in the summer and lower in the winter, as expected in the summer rainfall region of

South Africa.

Figure 6.22 Daily calibrated CRP soil moisture measurements in Catchment VI

6.4 Creating an In-situ Soil Moisture Dataset

There were three types/sets of in-situ soil moisture instrumentation used:

i. 5 TDR (pit) probes at one point placed horizontally at depths of 0.05, 010, 0.15, 0.20

and 0.30 m.

ii. 5 wireless TDR probes (0.12 m) placed vertically at different sites in the catchment.

iii. 8 echo probes placed horizontally at a depth of 0.10 m at different sites.

The TDR pit had the most complete data set out of all the in-situ instrument sets, as it has a

continuous record from the 11th of July 2014 to March 2015. The wireless TDR probes have

data from the 11th of July to March 2015; however, there are numerous gaps in the data set,

due to the erratic operation of all five TDR probes. The echo probes have the shortest data

record from the 9th of July 2014 to the 4th of September 2014. They were problematic and

were destroyed in a catchment fire on the 5th of September 2014.

0

10

20

30

40

50

60

3/1/2014 4/30/2014 6/29/2014 8/28/2014 10/27/2014 12/26/2014 2/24/2015

VW

C (

%)

Date

61

Although the study period of the research extends from March 2014 to March 2015, the

validation of the CRP was for the period 11th of July 2014 to the 20th of March 2015, as this

is the corresponding period for in-situ measurements.

In order to validate the CRP, a representative in-situ soil moisture data set is required. The

TDR pit, wireless TDR and ECHO probe datasets were converted from an hourly time-step to

a daily time-step, as a daily time-step was used in this research project.

The soil moisture from the TDR pit at five different depths in the profile is plotted against

time, as shown in Figure 6.23. The soil moisture at different depths followed the same

wetting and drying trends; however, the change in soil moisture, with depth was observed

(Figure 6.23). In general, the soil moisture increased with depth. In the “dry” period, the soil

moisture values were relatively constant. In the “wet” period, the temporal variation in soil

moisture was a result of more rainfall (input) and higher rates of total evaporation (output).

Figure 6.23 Daily TDR pit soil moisture measurements in Catchment VI

If we consider the effective measurement depth of the CRP during the one year period

between March 2014 and March 2015 (Figure 6.24), it can be seen that the effective

measurement depth of the CRP for this period does not exceed 0.15 m, therefore the 0.20 and

0.30 m TDR pit depth measurement data were excluded, when determining the average soil

moisture of the TDR pit. The 0.05 m TDR pit depth was also excluded, as the effective depth

of the CRP was closer to the 0.10 cm depth than the 0.05 m depth and the other two data

0

10

20

30

40

50

60

7/5/2014 10/3/2014 1/1/2015 4/1/2015

VW

C (

%)

Date

0.05 m 0.10 m 0.15 m 0.20 m 0.30 m

62

sources had measurements at 0.10 and 0.12 m depths (echo and wireless TDR). The mean

effective depth for this period was 0.11 m.

Figure 6.24 The CRP effective measurement depth

The average (0.10 and 0.15 m) daily TDR pit soil moisture was plotted against time (Figure

6.25). From Figure 6.25, it can be seen that the average TDR pit dataset covers the winter and

summer period.

Figure 6.25 The average TDR pit soil moisture measurements in Catchment VI

0.08

0.09

0.1

0.11

0.12

0.13

0.14

0.15

6/10/2014 7/30/2014 9/18/2014 11/7/2014 12/27/2014 2/15/2015 4/6/2015

De

pth

(m

)

Date

0

10

20

30

40

50

60

7/5/2014 8/24/2014 10/13/2014 12/2/2014 1/21/2015 3/12/2015

VW

C (

%)

Date

63

The daily wireless TDR data was plotted for the measurement period, as shown in Figure

6.26. The wireless TDR data has gaps in the data set. This is due to errors in the data logging

programme and faults in the sensors.

Figure 6.26 Daily wireless TDR data in Catchment VI

The Echo probe data was then plotted against time (Figure 6.27). The echo probe data only

covered the dry period, as the fire that burnt Catchment VI, also destroyed the sensors, which

could not be used again in this study.

Figure 6.27 Daily echo probe data in Catchment VI

0

10

20

30

40

50

60

7/5/2014 9/3/2014 11/2/2014 1/1/2015 3/2/2015

VW

C (

%)

Date

TDR1

TDR2

TDR3

TDR4

TDR5

0

10

20

30

40

50

60

7/5/2014 7/25/2014 8/14/2014 9/3/2014

VW

C (

%)

Date

ECHO1

ECHO2

ECHO3

ECHO4

ECHO5

ECHO6

64

All three (TDR pit, wireless TDR and echo probe) sets of data were merged by weighting

each point within the catchment equally, to create a representative in-situ soil moisture

estimate (Figure 6.28), which shows the merged dataset. This dataset weighted each sensor,

in each sample point equally.

Figure 6.28 Representative soil moisture measurements

6.5 Acquisition and Processing of the AMSR2 Soil Moisture Product

This section describes the methodology adopted to obtain the AMSR2 remote sensing soil

moisture product. The AMSR2 sensor estimates several parameters, which are predominately

linked to the energy and water cycles, namely, precipitation, water vapour, sea-surface

temperature, soil moisture and snow depth. The AMSR2 soil moisture product was obtained

from the JAXA website: http://sharaku.eorc.jaxa.jp/. The data is freely available; however, a

relatively short registration was required, in order to gain access into the GCOM-W1 Data

Providing Service.

Level Two and Level Three soil moisture data sets are available from July 2012 to present

day. The Level Two data is available on a 25 km spatial resolution grid, whilst the Level

Three data is available on both 10 km and 25 km spatial resolution grids. The Level Three

soil moisture product uses Level 1B brightness temperature and Level Two soil moisture data

and averages both of them spatially and temporally, with respect to predefined lattice grid

points on the earth’s surface.

0

10

20

30

40

50

60

7/5/2014 9/3/2014 11/2/2014 1/1/2015 3/2/2015

VW

C (

%)

Date

65

The 10 km and 25 km Level Three AMSR2 products are presented in Figure 6.29 and

illustrates the significance of spatial resolution. The 10 km spatial resolution is much finer

and therefore has a lot more detail, compared to the coarser 25 km resolution. The need for

finer spatial resolution data is to capture spatial heterogeinity and accurately represent

processes occuring at particular scales. The 10 km product results in a pixel area of 100 km2,

whilst the 25 km product results in a pixel area of 625 km2.

Figure 6.29 Comparison of spatial resolution of AMSR2 10 km and 25 km Level

Three soil moisture products

The Level Three soil moisture product at a 10 km spatial resolution was selected to be used in

this study. Once the Level Three 10 km soil moisture product is selected, the data type and

observation period needs to be specified, before specifying the search area. Once all the

specifications have been made, the data can be searched for and subsequently downloaded.

The data format is then specified and there is an option to leave the data in its original format

or convert it to HDF5, Geotiff of NetCDF. The conversion to Geotiff was chosen, as this data

format is compatible with ArcGis 9.3.

AMSR2 Level Three, 10km AMSR2 Level Three, 25km

66

The data can now be ordered. Once ordered, an order acceptance email, which contains an ftp

and URL site link to directly download the data. The data was then opened in ArcGis 9.3.

The data was defined on the WGS 84 geographic coordinate system and no projection or

transformation was required. There were two files per day, as each day had an ascending and

descending orbit.

Due to the nature of satellite orbits, on some days, the catchment area was covered by:

i. An ascending orbit;

ii. A descending orbit;

iii. An ascending and descending orbit;

iv. Neither.

The catchment shape file was overlaid onto the AMSR2 soil moisture raster image, as seen in

Figure 6.30. The pixel value that was covered by the catchment was determined. The pixel

values ranged from 0 to 600. The pixel values were multiplied by a scaling factor of 0.001, to

obtain the soil moisture in g/cm3. The density of water is assumed to be 1.0 g/cm3 and the

volume of 1.0 g of water is 1.0 cm3. Therefore, the VWC in percentage was obtained by

multiplying the scaled pixel value by 100.

Figure 6.30 The Catchment VI shapefile overlaid over the AMSR2 soil moisture

product to obtain the volumetric water content

Catchment VI

Pixel Value

67

6.6 Acquisition and Processing of the SMOS Soil Moisture Product

This section describes the methodology adopted to obtain the SMOS remote sensing soil

moisture product. Level Two and Level Three SMOS soil moisture products are available.

The Level Two soil moisture products are at a 30-50 km spatial resolution and can be

obtained from the ESA website: http://eopi.esa.int/. The data is freely available from the Eoli-

sa catalogue. The Eoli-sa catalogue gives users access to European Space Agency (ESA)

Earth Observation data products.

The Level Three SMOS soil moisture products are available from the national French and

Spanish processing entities, namely, the CATDS (Centre Aval de Traitement des Donnees

SMOS) and SMOS CP34-BEC (SMOS Barcelona Expert Centre). The BEC-34 was used, as

it was more user-friendly and the site was a lot easier to navigate. The BEC is an ESA expert

support laboratory committed to the ongoing analysis and development of new algorithms to

improve the baseline SMOS Level Two products.

The SMOS Level Three soil moisture products are computed from Level Two soil moisture

user data products, which consist of geophysical parameters, a theoretical estimate of

accuracy, flags and product descriptors. First, the Level Two soil moisture user data products

are filtered and combined into maps with the same spatial resolution. Next, the quality flags

and product descriptors are used to discard unreliable soil moisture values. The final soil

moisture product is a global daily soil moisture map on a 25 km spatial resolution grid.

The Level Three SMOS BEC-34 data was obtained from http://cp34-bec.cmima.csic.es/. This

site is the SMOS-BEC data distribution and visualization service. A short registration

procedure is required before the data can be accessed from the site’s THREDDS service,

which provides netCDF data files. The SMOS soil moisture products fall under the Land

category. The BEC land near real-time (0.25 x 0.25 degree resolution) was selected, as the

other two options had fine resolution soil moisture, which only covered the Iberian Peninsula.

The option selected was a global dataset. There are numerous soil moisture products at

different intervals. The one-day global soil moisture product was chosen. For each day, there

are two files, the ascending and descending orbits. Figure 6.31 shows the navigation through

the THREDDS service.

68

Figure 6.31 Navigation through the THREDDS service

The soil moisture product files can either be directly downloaded or viewed, using one of the

data viewers on the site. The viewers include NetCDF-Java ToolsUI, Integrated Data Viewer

and the Godiva2 online tool. The Godiva2 online visualization tool was used to obtain the

soil moisture data (Figure 7.32).

69

Figure 6.32 Layout of the Godiva2 online visualization tool

The Godiva2 visualization tool is very user-friendly and the data was easily obtained. The

required data set was viewed by selecting the necessary data requirements (1). The data

requirements were selected as follows; the BEC research products land, near-real time global

maps, daily one-day maps, choose between ascending and descending (both are required, but

are obtained separately) and lastly soil moisture (SM).

Next, the date was specified (2). The data for a specific day is viewed by clicking on the date

on the calendar. The transition from day-to-day is quick.

The global map was used to locate the co-ordinates of the area required. The correct pixel

was located, which corresponded to the co-ordinates of the catchment. The pixel information,

which consisted of its longitude, latitude and soil moisture value, was displayed (3).

A check was done to determine if the catchment was contained in one pixel. This was done

by opening the image in Google Earth (4).

1

2

3

4

70

Once it was determined that the catchment was completely contained in one pixel (Figure

6.33), the soil moisture data could be obtained by changing the date and obtaining the data by

clicking on the corresponding pixel (Figure 6.34).

Figure 6.33 The position of the catchment in relation to a single pixel of the SMOS

soil moisture product

71

Figure 6.34 Godiva2 visualization tool showing the pixel value for the Catchment VI

This method of obtaining daily soil moisture data was efficient, as the data did not have to be

downloaded, stored, opened with the appropriate software, spatially defined and projected,

before overlaying the catchment shapefile and determining the soil moisture estimate. This

online viewer was more than adequate to obtain the daily soil moisture data for both

ascending and descending data sets for a one year period. The images obtained were near-real

time, as the present day data was available the following day. The pixel values were in

volumetric water content, so no conversion was required.

6.7 Analysis of AMSR2 and SMOS Remote Sensing Data

The AMSR2 and SMOS products are daily soil moisture products; however, there were days

in which the study area was not covered by an ascending or descending orbit (missing data).

Of the 365 day study period the various days in terms of coverage are presented in Figures

6.35 and 6.36.

From Figure 6.35, it can be seen that AMSR2 data was available for 343 days of the 365

days. There were 22 days which had no data, resulting in a daily annual coverage of 94%.

The days with missing data seemed to follow a general pattern throughout the year, as there

were similar amounts of missing data each month of the observation period.

72

Figure 6.35 AMSR2 data availability for the observation period

From Figure 6.36, it can be seen that of the 365 of days observed SMOS data, 129 days are

missing, which results in a daily annual coverage of 65%. Of the total 236 days with data,

only 103 of them were covered by both ascending and descending orbits, whilst 66 days had

an ascending value only and 67 days had a descending value only.

Figure 6.36 SMOS data availability for the observation period

The 160 days for which ascending and descending AMSR2 data was available was plotted

against time, as indicated in Figure 6.37. This was done to compare the ascending and

descending values of the same day. From figure 6.37, there is very close correlation between

the ascending and descending values during the dry period. There is less correlation during

the wet periods. Both the ascending and descending datasets follow the same trend.

91

92

160

22

Ascending

Descending

Both

Missing

66

67

103

129Ascending

Descending

Both

Missing

73

Figure 6.37 Time series analysis of days which have both ascending and descending

AMSR2 values

The 103 days for which ascending and descending SMOS data was available was plotted as a

time series (Figure 6.38). From figure 6.38, it is evident that both the ascending and

descending data follow the same trend. There are differences in values between the two

datasets on the same day. These differences are larger in the wetter periods. The differences

in ascending and descending values are greater for the SMOS dataset compared to the

AMSR2 dataset.

Figure 6.38 Time series analysis of days which have both ascending and descending

SMOS values.

0

5

10

15

20

25

30

35

40

45

50

3/1/2014 4/20/2014 6/9/2014 7/29/2014 9/17/2014 11/6/2014 12/26/2014 2/14/2015 4/5/2015

VW

C (

%)

Date

Ascending Descending

0

10

20

30

40

50

60

70

3/1/2014 4/30/2014 6/29/2014 8/28/2014 10/27/2014 12/26/2014 2/24/2015

VW

C (

%)

Date

Ascending Descending

74

The differences observed between the ascending and descending values for the same day and

over the time period can be attributed to the temporal variation in soil moisture. The

ascending values are obtained at 06:00, whereas the descending values are obtained at 18:00.

Therefore, the change in soil moisture during this 12-hour interval could be due to rainfall

and total evaporation.

This is further emphasized in the dry period, when rainfall and total evaporation is low. The

ascending and descending values correlate well, as the VWC does not change much in the

interval. However, in summer, the rainfall and total evaporation rates are high resulting in

changes in the VWC between the ascending and descending measurements.

There are expected differences between ascending and descending overpass soil moisture

retrievals, due to the difference in geophysical conditions at day time and night time, with

night time conditions being more favourable for soil moisture retrieval (Kim et al., 2015).

Therefore, should soil moisture studies only use descending values, which would result in

more missing data days, but better favourable soil moisture retrieval values.

An alternative is to use only the days which have both ascending and descending values, as

this will average the two retrieval values, however, there will also be a large number of

missing data days. For this research study, all observation days with ascending, descending

and days with both ascending and descending values were used.

6.8 PyTOPKAPI Soil Moisture Product (SAHG)

This section details the acquisition of the PyTOPKAPI modeled soil moisture estimates. The

PyTOPKAPI soil moisture product was obtained from the Satellite Applications and

Hydrology Group (SAHG) website: http://sahg.ukzn.ac.za/.

The PyTOPKAPI modelled soil moisture product will be referred to as the SAHG soil

moisture product. The soil moisture products are freely and readily available. The data has a

temporal resolution of three hours and is on a spatial resolution grid ≈ 12.5 km. The data

ranges from the year 2008 to the present.

75

The data was obtained by downloading the SSI (soil saturation index) product in the ascii file

format, which can be opened and processed in ArcGis 9, as seen in Figure 6.39. The soil

moisture data for each day was available on a three-hour time-step. Therefore, eight files had

to be downloaded for each day.

Figure 6.39 Navigation through the SAHG website to download SSI data

The files were then opened in ArcGis 9.3 and the spatial referencing of the data was first

checked. The data was defined in the correct spatial reference (WGS 84 geographic co-

ordinates) and no defining projections and transformations were required. The soil moisture

products were on a three-hour time-step: (i) 00:00, (ii) 03:00, (iii) 06:00, (iv) 09:00, (v)

12:00, (vi) 15:00, (vii) 18:00 and (viii) 21:00 (see Figure 6.40). These eight raster images per

day had to be converted into a daily raster image. This was done by using the Mosaic to New

Raster option under the Data Management Tools in ArcGIS 9.3.

76

Figure 6.40 One day of SSI data (eight images on a three hour interval)

Once the daily image was created (Figure 6.41), the catchment shape file was overlaid onto it

(Figure 6.42). The pixel value was then determined and recorded.

77

Figure 6.41 Daily SAHG soil moisture product

Figure 6.42 Overlay of the Cathedral Peak Catchment VI onto the SAHG soil

moisture product

78

The soil moisture data from the SAHG product is represented as a soil saturation index (SSI),

which is expressed as a percentage. The SSI was then converted into a VWC, using the

simplified equation.

𝑆𝑆𝐼 = 𝑉𝑊𝐶

𝑃𝑜𝑟𝑜𝑠𝑖𝑡𝑦 (6.15)

The SSI equation (Equation 6.15) can easily be rearranged to solve for VWC:

𝑉𝑊𝐶 = 𝑆𝑆𝐼 × 𝑃𝑜𝑟𝑜𝑠𝑖𝑡𝑦 (6.16)

The soil porosity can be determined by measurement or calculation. In order to calculate

porosity, the following equation was used:

𝑃𝑜𝑟𝑜𝑠𝑖𝑡𝑦 = 1 − 𝑃𝑏

𝑃𝑠 (6.17)

where Pb is the bulk density and Ps is the particle density. The bulk density was determined

to be 0.593 g/cm3 and a generic value of 2.65 g/cm3 was used for the particle density (Hillel,

2008).

𝑃𝑜𝑟𝑜𝑠𝑖𝑡𝑦 = 1 − 0.593

2.65

The porosity was calculated to be 0.77. Therefore, from Equation 6.16, all the SSI values

were multiplied by 0.77, to convert them to volumetric water content.

The porosity was determined from the bulk density, which was valid for the top 0.15 m of the

soil surface. The PyTOPKAPI soil moisture product is the SSI for the A and B horizon of the

soil profile. Therefore, a porosity value that represents the A and B (one meter depth)

horizon, needs to be determined. A study conducted by Everson et al. (1998), contains soil

bulk density values for Catchment VI at depths of 0.20, 0.50, 0.75 and 1.00 m (Table 6.6).

The bulk density values used in this study can be incorporated and used to obtain a

representative porosity value that covers the A and B horizons. The porosity values should

not have changed over the past 16 years, as the catchment is in a protected area.

79

Table 6.6 The bulk density of Catchment VI, as determined by Everson et al. (1998)

Soil Depth (m) Dry Season Wet Season Mean 0.20 0.660 0.786 0.868 0.886 0.800 0.50 0.898 0.899 0.847 0.798 0.861 0.75 0.803 0.757 0.662 0.784 0.752 1.00 0.863 0.799 0.905 0.844 0.853

These bulk density values were incorporated into the values that were obtained in this study

and the mean bulk density of the soil profile (0 to 1.00 m) was determined (Table 6.7).

Table 6.7 Mean bulk density according to depth

Soil Depth (m) Mean Bulk density (g/cm3) 0.05 0.525 0.10 0.597 0.15 0.656 0.20 0.800 0.50 0.861 0.75 0.752 1.00 0.853 Average 0.720

The average bulk density was 0.720 g/cm3. The average porosity was calculated, using the

new bulk density value in Equation 6.12.

𝑃𝑜𝑟𝑜𝑠𝑖𝑡𝑦 = 1 − 0.720

2.65 (6.12)

The average porosity is 0.728. Therefore, the SSI pixel values obtained must be multiplied by

0.728, in order to convert them into VWC (Equation 6.11).

80

6.9 Surface Energy Balance System (SEBS)

The SEBS Model, developed by Su (2002), was run in the Integrated Land and Water

Information System (ILWIS) 3.8.3 for the estimation of relative evaporation and evaporative

fraction in Cathedral Peak Research Catchment VI for the period of March 2014 to March

2015. The SEBS Model uses remote sensing and meteorological data sets to estimate heat

fluxes.

Landsat 8 images were used for the estimation of relative evaporation and evaporative

fraction, using the SEBS Model. The Landsat 8 satellite, as seen in Figure 6.43, is the latest

addition to the Landsat series and was launched on the 11th of February 2013 (Markham et

al., 2015). Landsat satellites have continuously acquired information of the earth’s land

surface since 1972, thus the continuation of data acquisition from the Landsat 8 satellite is

essential (USGS, 2013). Landsat 8 orbits the earth at an altitude of 705 km, which results in

14 full orbits being completed each day, with every point of the earth being covered once

every 16 days. The satellite carries out north to south orbits and has an overpass time of 10:00

(Markham et al., 2015). Landsat 8 carries two push-broom sensors, so that the reflective and

thermal bands were split into two instruments (Markham et al., 2015). These two instruments

are the Operational Land Imager (OLI) and Thermal Infrared Sensor (TIRS).

Figure 6.43 Landsat 8 satellite

81

Landsat 8 images were used in this study due to their spatial resolution. It has the ability to

estimate relative evaporation and evaporative fraction at a spatial resolution of 30 m,

however, it has a temporal resolution of 16 days. The Landsat 8 product consists of several

bands (one to eleven), each band having its own spectral characteristics and wavelength range

(Table 6.8).

Table 6.8 Characteristics of the various landsat 8 bands (USGS, 2015).

Spectral bands Wavelength (μm) Resolution (m) 1- coastal/aerosol 0.43-0.45 30 2- blue 0.45-0.51 30 3- green 0.53-0.59 30 4- red 0.64-0.67 30 5- near IR 0.85-0.88 30 6- SWIR 1 1.57-1.65 30 7- SWIR-1 2.11-2.29 30 8- panchromatic 0.50-0.68 15 9- cirrus 1.36-1.38 30 10- TIRS 1 10.60-11.19 100 11- TIRS 2 11.50-12.51 100

Landsat 8 satellite imagery was freely acquired from the Earth Explorer website:

http://earthexplorer.usgs.gov. The area of interest and time period of the data required was

specified. Next, the data set that the user requires can be selected from a list of all the

available data sets. The data set selected was the L8 OLI/TIRS. The catchment study area fell

within one Landsat 8 image. The Landsat 8 image footprint is very large, as seen in Figure

6.44. Between March 2014 and March 2015, there were a total of 22 images.

Figure 6.44 Landsat 8 image footprint

82

Most of the images, which were acquired, were for clear sky conditions. Some images

possessed variable cloud cover; however, the area of interest was cloud-free, therefore, the

image could still be used. There were six images, which could not be used for the retrieval of

the evaporative fraction and relative evaporation, as the image contained cloud cover, which

blocked the area of interest. The different cloud-cover conditions are seen in Figure 6.45.

Figure 6.45 Landsat images with different cloud-cover conditions

Of the 22 images available for the study period, only 16 images could be used due to cloud-

cover. The 16 images that could be used were downloaded. The Level one Geotiff Data

Product was the download option selected. Each Level One Geotiff product contains a

Material Library (MTL) file, which contains all the necessary information (metadata) about

the images and 12 images, which include images B1 to B11 and BQA, where B1 is band one

and B11 is band eleven.

83

The ILWIS 3.8.3 software was downloaded from http://52north.org/downloads/ilwis/. The

images (bands) were imported into ILWIS 3.8.3 as digital numbers (DN). The bands required

were 2, 3, 4, 5, 6, 7, 10 and 11. These bands were rescaled to Top of Atmosphere (TOA)

reflectance and/or radiance, using the radiometric rescaling coefficients provided in the

metadata file. The procedure followed has been outlined by Allen et al. (2002) and the USGS

(2015). The USGS (2015) provides the necessary equations for the conversion to TOA

Radiance, conversion to TOA Reflectance and conversion to TOA brightness temperature.

The equations to obtain the inputs into the SEBS Model, such as the albedo, NDVI, surface

emissivity and land surface temperature maps are presented in Allen et al. (2002). Although

the equations in Allen et al. (2002) were intended to be used in SEBAL, they can be used in

other models, such as SEBS.

Conversion to TOA Radiance

Bands 10 and 11 (TIRS bands) were converted to TOA Radiance, using the radiance

rescaling factors provided in the metadata file (USGS, 2015).

Lλ = (𝑀𝐿 ∗ 𝑄𝑐𝑎𝑙) + 𝐴𝐿 (6.18)

where, Lλ is the TOA spectral radiance (Watts/(m2*srad*µm)), ML is the band specific

multiplicative rescaling factor (RADIANCE_MULT_BAND_X) from the metadata file

(where X is the band number), Qcal is the quantized and calibrated standard product pixel

value (DN) and AL is the band specific additive rescaling factor

(RADIANCE_ADD_BAND_X) from the metadata (where X is the band number).

Conversion to TOA Reflectance

Bands 2, 3, 4, 5, 6 and 7 (OLI bands) were converted to TOA planetary reflectance, using

reflectance rescaling coefficients provided in the metadata file (USGS, 2015).

Pλ′ = (𝑀𝑃 ∗ 𝑄𝑐𝑎𝑙) + 𝐴𝑃 (6.19)

84

where Pλ′ is the TOA planetary reflectance (without solar angle correction), Mp is the band

specific multiplicative reflectance factor (REFLECTANCE_MULT_BAND_X) from the

metadata file (where X is the band number) and Ap is the band specific additive rescaling

factor (REFLECTANCE_ADD_BAND_X) from the metadata (where X is the band number).

The TOA reflectance was then corrected for the sun angle (USGS, 2015):

Pλ =Pλ′

sin (Θ𝑆𝐸) (6.20)

where pλ is the TOA planetary reflectance and ΘSE is the local sun elevation

(SUN_ELEVATION), which is obtained from the metadata file.

The ESUN values were determined for bands 2, 3, 4, 5, 6 and 7, using the following

equations (Allen et al., 2002):

𝐸𝑆𝑈𝑁 = (𝜋 ∗ 𝑑2) ∗ (𝑅𝐴𝐷_𝑀𝐴𝑋

𝑅𝐸𝐹_𝑀𝐴𝑋) (6.21)

Where d is the earth-sun distance, RAD_MAX is the maximum radiance and REF_MAX is

the maximum reflectance (all of which are found in the metadata file).

The ESUN values (Table 6.9) are required to create an equation to determine albedo for the

TOA (Allen et al., 2002).

𝛼𝑇𝑂𝐴= ∑(𝜔λ ∗ Pλ) (6.22)

𝜔λ =𝐸𝑆𝑈𝑁λ

∑𝐸𝑆𝑈𝑁λ (6.23)

Table 6.9 Calculated ESUN values

Band ωλ Band ωλ 2 0.300 5 0.143 3 0.277 6 0.036 4 0.233 7 0.012

85

The ESUN band values change, due to changes in the earth-sun distance (d), maximum

radiance and maximum reflectance of each data product. However, the 𝜔λ band value was

determined to be the same for each image, as the ratio remains the same.

Albedo is then calculated by correcting TOA albedo, using the following equation (Allen et

al., 2002):

∝=𝛼𝑇𝑂𝐴−𝛼𝑃𝑎𝑡ℎ_𝑟𝑎𝑑𝑖𝑎𝑛𝑐𝑒

𝜏𝑠𝑤2 (6.24)

where αPath_radiance is 0.03 and τsw is 0.774 (Allen et al., 2002).

The generated albedo map is illustrated in Figure 6.46. Albedo ranges from 0.0 to 1.0 and is a

measure of the reflectivity of the earth’s surface. As the reflectivity increases, the albedo

value increases. Grasslands generally have an albedo of between 0.15 to 0.25.

Figure 6.46 Albedo map generated in ILWIS

86

The normalized difference vegetation index (NDVI) was determined by Allen et al. (2002):

𝑁𝐷𝑉𝐼 =(𝑃5−𝑃4)

(𝑃5+𝑃4) (6.25)

where P5 is the corrected reflectance band five and P4 is the corrected reflectance band four.

The generated NDVI map is illustrated in Figure 6.47. The NDVI ranges from -1.0 to 1.0.

The more vegetation present, the higher the NDVI value. Water bodies have a negative NDVI

value.

Figure 6.47 NDVI map generated in ILWIS

The surface emissivity (εo) was determined using the following equation Allen et al. (2002):

𝜀𝑜 = 1.009 + 0.047 ∗ ln (𝑁𝐷𝑉𝐼) (6.26)

The generated surface emissivity map is illustrated in Figure 6.48. The surface emissivity is

0.999 for NDVI values less than 0.

87

Figure 6.48 Surface emissivity map generated in ILWIS

The TIRS bands (bands 10 and 11) are converted from spectral radiance to at-satellite

brightness temperature, using the following equation (USGS, 2015):

𝑇𝑏𝑏 = 𝐾2

𝑙𝑛[𝐾1

𝐿λ+1]

(6.27)

Where K1 and K2 are constants, which are found in the metadata file and Lλ is either band

10 or band 11, according to high or low gain conditions. Band 11 was used in this study, as

high gain is suited for grassland vegetation.

The Land Surface Temperature (LST) can be determined by Allen et al. (2002):

𝐿𝑆𝑇 = 𝑇𝑏𝑏

𝜀𝑜0.25 (6.28)

Figure 6.49 shows the generated land surface temperature map. The land surface temperature

is expressed in degrees kelvin.

88

Figure 6.49 Land surface temperature map generated in ILWIS

The SEBS Model requires a digital elevation model (DEM) (Figure 6.50), as input. The DEM

of South Africa was obtained http://srtm.csi.cgiar.org. The DEM for each image had to be

resampled before it could be used:

Figure 6.50 DEM map used in ILWIS as an input in SEBS

89

Once the input maps (albedo, NDVI, surface emissivity, land surface temperature and DEM)

had been created, the SEBS Model was run in ILWIS. The meteorological data required to

run the model was obtained from the eddy covariance weather data, which is located within

Catchment VI. On the days where the eddy covariance weather data was not available, data

from the nearby Mikes Pass weather station was used. The data required from the weather

station were the instantaneous downward solar radiations, wind speed, air temperature,

pressure, mean daily air temperature and sunshine hours. The downward solar radiation, wind

speed, air temperature and pressure data were at the time of the satellite overpass (10:00 am).

Figure 6.51 The SEBS model in ILWIS

The SEBS Model was run (Figure 6.51) and many outputs were generated. These outputs

included daily evaporation, relative evaporation, evaporative fraction, net radiation, soil heat

flux, sensible heat flux dry, sensible heat flux wet, sensible heat flux index and leaf area

index.

90

For the purpose of this research, the evaporative fraction (Figure 6.52) and relative

evaporation (Figure 6.53) outputs were required.

Figure 6.52 Evaporative fraction map generated as an output of the SEBS model in

ILWIS

Figure 6.53 Relative evaporation map generated as an output of the SEBS model in

ILWIS

91

The relative evaporation output map (Figure 6.53) was exported as in ASCII format. The

output ASCII file was opened in ArcGIS 9.3 (Figure 6.54). The spatial reference of the

relative evaporation map was defined as WGS 1984 UTM Zone 35 N. The catchment

shapefile was projected to WGS 1984 UTM Zone 35 N and overlaid onto the relative

evaporation map (Figure 6.55).

Figure 6.54 Catchment VI shapefile overlaid onto relative evaporation map

Figure 6.55 The relative evaporation map of Catchment VI

Catchment VI

Catchment VI

92

The spatial analyst tool was then used in ArcGIS 9.3 data management tools, to calculate the

zonal statistics of the area enclosed by the catchment shapefile. Table 6.10 shows the zonal

statistics, it can be seen that the catchment encloses 686 pixels. The mean value is taken from

the zonal statistics as the relative evaporation for that image. The same procedure was

repeated to determine the evaporative fraction values.

Table 6.10 Zonal statistics of relative evaporation enclosed by catchment VI

Once the relative evaporation and evaporative fraction values had been determined, soil

moisture was calculated, using two equations, one by Su et al. (2003) (Equation 4.13) and the

other by Scott et al. (2003) (Equation 4.14).

The SEBS process is depicted in the flow diagram (Figure 6.56). It illustrates the various

procedures required to obtain relative evaporation and the evaporative fraction, using the

SEBS model.

93

Figure 6.56 Processes used to obtain relative evaporation from Landsat 8 data

Import

Bands 2

,3,4

,5,6

,7,1

0 a

nd 1

1

as D

igital

num

bers

(D

N)

in Ilw

is

Convers

ion to T

OA

Reflecata

nce

(Bands 2

,3,4

,5,6

and 7

)

Pλ′=

(𝑀𝑃∗𝑄

)+

𝐴𝑃

Sun a

ngle

corr

ection

Pλ=

Pλ′

sin(Θ

𝑆𝐸)

TO

A a

lbedo

𝛼𝑇 𝐴= ∑

(𝜔λ ∗ Pλ)

Alb

edo

∝=

𝛼𝑇 𝐴−

𝛼𝑃 𝑡

_𝑟 𝑑𝑖

𝑠

2

ND

VI

𝑁𝐷𝑉𝐼=

(𝑃5−

𝑃4)

(𝑃5+

𝑃4)

Convers

ion to T

OA

Radia

nce

(Bands 1

0 a

nd 1

1)

Lλ=

(𝑀

𝐿∗𝑄

)+

𝐴𝐿

Brightn

ess T

em

pera

ture

𝑇𝑏𝑏=

2

1

𝐿λ+

1

Land s

urf

ace tem

pera

ture

𝑇𝑆=

𝑇𝑏𝑏

𝜀𝑜0.25

Surf

ace e

mis

siv

ity

𝜀𝑜=

1.009+

0.047∗ln(𝑁

𝐷𝑉𝐼)

SEB

S

ILW

IS 3.8

.3

Mete

oro

logic

al D

ata

Sola

r ra

dia

tion

Win

dspeed

Tem

pera

ture

Pre

ssure

Sunshin

e h

ours

Avera

ge tem

pera

ture

DEM

Sun

zenith

angle

Julia

n d

ay

94

7. RESULTS AND DISCUSSION

This chapter presents the results and discussion of the various aspects of the research project.

The section will detail the following aspects:

i. The validation of the CRP with a representative in-situ soil moisture data set;

ii. The validation of remote sensing soil moisture products (AMSR2 and SMOS) with

CRP soil moisture estimates; and

iii. The validation of modelled soil moisture products (SAHG and SEBS) with CRP soil

moisture estimates.

The term validation is used in the sense of an assessment, rather than a strict justification to

observations. The validation will therefore focus on the comparison of spatial pattern,

temporal development and regularity (consistency of estimates produced) amongst the

different data sources. The types of analyses implemented in this research study include:

i. Time series analysis;

ii. Correlation coefficient (R) and coefficient of determination (R2);

a. The slope of the graph represents the change in y-units per change in x-units

(the change of variable two to the change in variable one). The closer the slope

value is to 1, the better the datasets correlate to one another;

b. R2 is the coefficient of determination and denotes the strength of the linear

association between variable x and variable y. R2 ranges from zero to one (0≤

R2 ≤ 1). The strength of the linear association increase as R2 becomes closer to

one;

iii. Residual analysis; and

iv. Paired t-test.

The statistical analyses selected for this study are amongst the most common analyses used in

current satellite-based soil moisture studies. These analyses are considered appropriate to

validate the various soil moisture products used in this study, in order to draw reliable

conclusions. The above statistical analyses, will not all be used on each validation, but will be

selected for use appropriately.

95

Soil moisture varies both spatially and temporally and fluxes in soil moisture content occur

over short time periods. The key input to soil moisture content is rainfall. Therefore, rainfall

data is necessary to describe and supplement changes in soil moisture. Rainfall data from a

rain gauge situated within Catchment VI was obtained for the one year period from March

2014 to March 2015 on an event basis. This data was then converted to a daily rainfall

dataset, as shown in Figure 7.1. The rainfall distribution shows that the majority of the

rainfall occurs in the summer months. During winter there are a few small rainfall events,

whilst in summer there are several rainfall events, which vary in magnitude.

Figure 7.1 Daily rainfall measured at Catchment VI

7.1 Validating the CRP

One of the key objectives of this research project is to test the suitability of the CRP for

providing spatial estimates of soil moisture. In order to achieve this, the CRP is validated

against a representative in-situ soil moisture data set.

A time series analysis was plotted to visually interpret and compare the CRP dataset and the

representative in-situ data set (Figure 7.2). Although the study period of the research project

was from the 1st of March 2014 to the 28th of February 2015 (one year), the time period in

which the representative data set was available (12th July 2014 to 28th February 2015) was

0

10

20

30

40

50

60

70

80

90

100

3/1/2014 5/1/2014 7/1/2014 9/1/2014 11/1/2014 1/1/2015

Rai

nfa

ll (m

m)

Date

96

used. This time period is considered adequate for the validation procedure, as both the wet

(summer) and dry (winter) periods are covered.

As shown in Figure 7.2, the CRP estimates followed the same seasonal trend as the in-situ

soil moisture data set. The CRP estimates correlated to the in-situ soil moisture data set better

in the wetter periods, when the soil moisture values were higher (above 30%), compared to

the drier periods. The CRP soil moisture estimates were higher than the in-situ estimates for

the dry period with low VWC values. This could be due to the differences in the

measurement depths between the CRP and the representative in-situ instruments. Overall, it

can be seen that the CRP dataset followed the general trend of the in-situ soil moisture

variations. The soil moisture fluctuations were dependent on the rainfall input and total

evaporation rate, as there were smaller fluctuations in winter, due to less rainfall and a lower

total evaporation rate. In summer, the fluctuations in soil moisture were greater, due to more

rainfall and a higher rate of total evaporation. There were differences between the datasets, as

the CRP dataset was generally slightly higher than the in-situ dataset. This difference was

more noticable when the soil moisture content was low. Overall, the CRP data seemed to

correlate well with the in-situ soil moisture dataset.

Figure 7.2 Daily in-situ and CRP soil moisture estimates for Catchment VI

A graph of correlation coefficients was then plotted with the representative in-situ data set on

the x-axis and the CRP estimates on the y-axis (Figure 7.3). From Figure 7.3, it can be seen

0

10

20

30

40

50

60

0

10

20

30

40

50

60

70

80

90

100

7/11/2014 9/11/2014 11/11/2014 1/11/2015

VW

C (

%)

Rai

nfa

ll (m

m)

Date

Rainfall In-situ CRP

97

that the majority of the points are above the 1:1 line (red), however they are in close

proximity to the 1:1 line. There seems to be no extreme outliers. The positive y-intercept of

12.024 indicates an over-estimation by the CRP of the lower values (when the soil is drier).

The slope is 0.709, which represents the relationship between the variables, with respect to

their increases and decreases. This indicates that there is a good correlation between both data

sets, as the value is close to 1. The R2 is 0.845. It is a measure of the strength of the linear

association between the datasets.

Figure 7.3 Scatterplot of In-situ soil moisture estimates against CRP estimates

A graph of the residuals was then plotted (Figure 7.4). The ∆ VWC (%) was the difference

between the in-situ (independent) and CRP (dependent) datasets. This graph is plotted to

illustrate how the difference in variables change over time. From Figure 7.4, it can be seen

that the majority of the residuals are negative, which indicates that the CRP soil moisture

values are higher than those of the subsequent representative in-situ values. When the soil

moisture content is high in the wet period, the residuals are lower in absolute value, which

indicates less of a difference between the data sets. The residuals are negative in the dry

periods and they also have the highest absolute value. In the wet periods, the residuals are

lower and are mostly negative; however, there are some positive values.

y = 0.7099x + 12.024R² = 0.8445

0

10

20

30

40

50

0 10 20 30 40 50

CR

P (

WC

(%

)

In-situ VWC (%)

98

Figure 7.4 Residual graph of In-situ against CRP soil moisture estimates

A paired T-test was then conducted on the two data sets (Table 7.1). This analysis indicates

whether the datasets were significantly different to each other. From Table 7.1, it can be seen

that the mean of the CRP is greater than that of the in-situ dataset, which results in the t-stat

value being negative. The variance, which is the average of the squared differences from the

mean, is the standard deviation squared. There are 436 degrees of freedom (df). The absolute

value of the t-stat (4.827) is greater than the critical two-tail value (1.965) and the P value

(1.919x10-6) is less than the alpha value (0.05). Therefore, there is a significant difference

between the datasets.

Table 7.1 T-test of In-situ against CRP estimates

In-situ CRP Mean 29.557 33.005 Variance 74.464 44.433 Observations 233 233 Hypothesized Mean Difference 0 df 436 t Stat -4.827 P(T<=t) two-tail 1.919x10-6 t Critical two-tail 1.965

-12

-8

-4

0

4

8

12

7/11/2014 9/11/2014 11/11/2014 1/11/2015

∆ V

WC

(%)

Date

99

The inconsistencies seen in the data comparison could be due to the following. The

calibration and validation of the CRP is not a straight-forward task. The calibration of the

CRP could contain bias errors, due to the sampling points selected. If different sampling

points were chosen, the soil moisture estimates could be different.

The in-situ dataset that was used to validate the CRP was not the actual average of the

catchment, but the average of the sensors in the catchment. If the sensors were placed in

different locations in the catchment, the in-situ dataset could be different. The area covered

by the in-situ sensors was far smaller than that covered by the CRP. The in-situ soil moisture

sensors were a maximum of 200 m away from the CRP, however, the CRP has a

measurement footprint that exceeds 200 m in radius. In addition, the representative data set

was biased towards the TDR pit as it was the only continous dataset, which resulted in one

single point weighing more than any other point.

The CRP was not measuring soil moisture at a constant depth. The CRP measurement depth

was dependent on the soil moisture content. Thus, when the soil was dry, the CRP probe was

measuring at a deeper depth than when the soil was wet. The representative in-situ data set

was meassuring constantly at an average depth of about 12 cm, whilst the CRP was

measuring at around 14 cm when the soil is dry and 11 cm when the soil was wet. Therefore,

the dry periods did not correlate as well as the wet periods, as the measurement depths were

different. Overall, it can be seen that the CRP was suitable to provide spatial estimates of soil

moisture. The CRP probe data set will then be used to validate the satellite-based estimates of

soil moisture.

7.2 AMSR2 Soil Moisture Product Validation

The AMSR2 Level Three soil moisture product is on a 10 km spatial resolution grid.

Although this grid is relatively small in comparison to other remote sensing soil moisture

products, it is still very large in comparison to the catchment area. The catchment area is 0.68

km2, whilst the pixel area is 100 km2. Therefore the pixel is 147 times larger than the study

area. However, validating remote sensing soil moisture products, such as the AMSR2

product, with CRP estimates is an improvement from validating remote sensing soil moisture

products with in-situ point measurements. For this analysis, the one-year study period of the

1st of March 2014 to the 28th of February 2015 was used. A time series analysis graph was

100

plotted to illustrate the characteristics of the AMSR2, CRP and rainfall data over time (Figure

7.5).

From Figure 7.5, it can be seen that the AMSR2 soil moisture product underestimated soil

moisture throughout the study period, when compared to the CRP. The AMSR2 soil moisture

product followed the seasonal trend of the CRP estimates. The discontinuation in the

AMSR2 line was due to missing data. The AMSR2 dataset fluctuated more in the wet

periods, whilst it fluctuated less in the dry periods. Although the AMSR2 data set

underestimated soil moisture, it followed a similar trend in daily soil moisture fluctuation and

seemed to correlate well with the CRP data set.

Figure 7.5 Time series analysis of CRP and AMSR2 soil moisture estimates

A day in summer and winter that had both an ascending and descending value, are shown in

Figure 7.6. In summer, the spatial and temporal variation of the pixel values varied more than

in winter. This was seen by the changes in the pixel values between the ascending and

descending values. The differences between summer ascending and descending pixel values

were greater than the differences between winter ascending and descending values. This

showed the heterogenous nature of soil moisture, as well as the difference in soil moisture

dynamics due to seasonality.

0

10

20

30

40

50

60

0

10

20

30

40

50

60

70

80

90

100

3/1/2014 5/1/2014 7/1/2014 9/1/2014 11/1/2014 1/1/2015

VW

C (

%)

Rai

nfa

ll (m

m)

Date

Rainfall CRP AMSR2

101

Figure 7.6 Comparison between summer and winter images of AMSR2 soil moisture

estimates

102

A scattergraph of CRP (x-axis ) against AMSR2 (y-axis) was then plotted (Figure 7.7). From

Figure 7.7, all the data points are below the 1:1 line (red), which shows that the AMSR2

dataset is always under-estimating soil moisture in comparison to the CRP. The slope of the

graph is 0.649, which is good. There is a positive correlation and the R2 is 0.505, which

indicates a good fit. There are some outliers above and below the line of best fit.

Figure 7.7 Scatterplot of CRP against AMSR2 soil moisture estimates

A residual graph of the CRP against AMSR2 soil moisture was then plotted against time

(Figure 7.8). The ∆ VWC (%) was the difference between the CRP and AMSR2 datasets.

From the residual graph (Figure 7.8), it can be seen that the AMSR2 dataset was under-

estimating soil moisture throughout the study period. In the dry period, the residuals were

fairly constant, whilst in the wet periods, the residuals fluctuated more. The mean residual

was 22.81. The residuals were constantly lower in the dry period, compared to the wet period,

which indicates that there was less underestimation of the AMSR2 data in the dry period.

y = 0.6492x - 11.125R² = 0.505

0

20

40

60

0 20 40 60

AM

SR2

VW

C (

%)

CRP VWC (%)

103

Figure 7.8 Residual graph of CRP against AMSR2 soil moisture estimates

A paired t-test was then conducted (Table 7.2). In order to properly perform a t-test, paired

values were required. Therefore, the CRP values corresponding to the AMSR2 days, which

had no data, were removed. From Table 7.2, it can be seen that the means are very different,

as the mean of the CRP is 33.326 and AMSR2 is 10.511. The variance values are similar, as

the variance of the CRP is 50.010 and the AMSR2 is 41.737. There are 669 degrees of

freedom. The t-stat value (43.791) exceeds the t critical value (1.964) and the P value

(1.3x10-198) is less than alpha (0.05). Therefore, there is a significant difference between the

datasets.

Table 7.2 T-test of CRP against AMSR2 estimates

CRP AMSR2 Mean 33.326 10.511 Variance 50.010 41.737 Observations 338 338 Hypothesized Mean Difference 0 df 669 t Stat 43.791 P(T<=t) two-tail 1.3x10-198 t Critical two-tail 1.964

When the AMSR2 product data was processed and the pixel values obtained, it was found

that the daily data values corresponded to four different situations. There were days when

only a descending pixel value was available (10th of December 2014) and there were days

when only an ascending pixel value was available (11th of December 2014) (Figure 7.9).

0

5

10

15

20

25

30

35

40

45

50

3/1/2014 5/1/2014 7/1/2014 9/1/2014 11/1/2014 1/1/2015

∆ V

WC

(%)

Date

104

There were days when both an ascending and descending pixel value was available (14th of

December 2014) and there were days when no pixel data was available for the catchment area

(15th of December 2014) (Figure 7.10).

Figure 7.9 A day with a descending and a day with an ascending value for the AMSR2

soil moisture product

105

Figure 7.10 A day with both an ascending and descending value, and a day with no

value for the AMSR2 soil moisture product

106

7.3 SMOS Soil Moisture Product Validation

The SMOS Level Three soil moisture product is on a 25 km spatial grid. Although this grid is

smaller than that of the Level Two product (40 km), it is still very large in comparison to the

catchment area. The catchment area is 0.68km2, whereas the pixel area is 625km2. Therefore,

the pixel is 920 times larger than the study area.

A graph of the time series analysis of the daily CRP and SMOS datasets were plotted (Figure

7.11). The time period used, was a one year period from the 1st of March 2014 to the 28th of

February 2015. The SMOS soil moisture product is a daily product; however, the satellite

coverage does not scan the entire earth’s surface in one-day. The data set has a lot of gaps, as

only 236 images during the study period have data for the study area. Due to the numerous

gaps in the SMOS dataset, the data was plotted as points to improve the representation of the

data. From Figure 7.11, it can be seen that the SMOS soil moisture estimates follow the same

general trend of the CRP. The SMOS dataset underestimates soil moisture the majority of the

time, when compared to the CRP. There are times in the wet periods, when the SMOS

product over-estimates soil moisture. In the wet periods, the fluctuation in soil moisture

estimates of SMOS is great. This fluctuation is less in the dry periods, due to greater fluxes in

soil moisture during summer when compared to winter.

Figure 7.11 Time series analysis of CRP and SMOS soil moisture estimates for

Catchment VI

0

10

20

30

40

50

60

70

0

10

20

30

40

50

60

70

80

90

100

3/1/2014 5/1/2014 7/1/2014 9/1/2014 11/1/2014 1/1/2015

VW

C (

%)

Rai

nfa

ll (m

m)

Date

Rainfall SMOS CRP

107

A day in summer and winter with both ascending and descending SMOS data, are shown in

Figures 7.12, 7.13, 7.14 and 7.15. The summer day image showed that the soil moisture was

relatively higher on the east coast of South Africa and that soil moisture was high in the

mountainous areas (Figures 7.12 and 7.13).

Figure 7.12 Ascending SMOS image in summer (17 December 2014)

Figure 7.13 Descending SMOS image in summer (17 December 2014)

The day in winter shows that soil moisture is generally low in South Africa (Figures 7.14 and

7.15). The areas with the higher soil moisture values are also the areas that receive the most

rainfall throughout the year.

108

Figure 7.14 Ascending SMOS image in winter (15 August 2014)

Figure 7.15 Descending SMOS image in winter (15 August 2014)

A graph of CRP (x-axis) was plotted against SMOS (y-axis) (Figure 7.16). From Figure 7.16,

it can be seen that the majority of the points lie below the 1:1 line (red). This indicates that

the SMOS data set generally under-estimates soil moisture. There are some points on the 1:1

line and several points above the 1:1 line. The R2 value is 0.485, which indicates a good

correlation. The slope of the graph is 1.2797.

109

Figure 7.16 Scatterplot of CRP against SMOS soil moisture estimates

A graph of the residuals against time was plotted (Figure 7.17). The ∆ VWC (%) is the

difference between the CRP and SMOS datasets. The spaces between the columns are where

there is missing SMOS data. From the Figure 7.17, it is seen that the residuals range from -23

to +30. The residuals are positive, except for a few instances in the wet periods, when the

SMOS estimates are higher than the CRP measurements. The residuals fluctuate more than

the AMSR2 soil moisture residuals.

Figure 7.17 Residual graph of CRP against SMOS soil moisture estimates

y = 1.2797x - 20.506R² = 0.4853

0

10

20

30

40

50

60

70

0 10 20 30 40 50 60 70

SMO

S V

WC

(%

)

CRP VWC (%)

-30

-20

-10

0

10

20

30

3/1/2014 5/1/2014 7/1/2014 9/1/2014 11/1/2014 1/1/2015

∆ V

WC

(%)

Date

110

A paired t-test was conducted (see Table 7.3). The t-test requires pairs of x and y values, thus

the CRP data, which had no subsequent SMOS data, had to be removed. From table 7.3, it

can be seen that the mean of the CRP dataset (33.396) is higher than that of the SMOS

dataset (22.231). The variance of the SMOS dataset is very large (166.267), compared to that

of the CRP dataset (49.276). There are 363 degrees of freedom. The t stat value is 11.683,

which is larger than the critical two-tail value of 1.967 and the p value is 5.49x10-27, which is

smaller than the alpha value (0.05). Therefore, there is a significant difference between the

datasets.

Table 7.3 T-test of CRP against SMOS estimates

CRP SMOS Mean 33.396 22.231 Variance 49.276 166.267 Observations 236 236 Hypothesized Mean Difference 0 df 363 t Stat 11.683 P(T<=t) two-tail 5.49x10-27 t Critical two-tail 1.967

Like the AMSR2 dataset, there were four different situations of days processed. These

included ascending only, descending only, both (ascending and descending) and neither.

The SMOS data set had cases where there was an ascending and/or descending band covering

the area; however, there were pixels missing within the band itself, which resulted in the area

having no pixel value (Figure 7.18).

Figure 7.18 SMOS missing data within band

111

7.4 Comparing Remote Sensing Soil Moisture Products

In order to compare the two remote sensing soil moisture products, both were plotted on a

time series graph along with CRP estimates (Figure 7.19). The AMSR2 and CRP were

plotted on a line graph, whilst the SMOS data was plotted on scatter points due to the large

number of missing data. It is difficult to compare the remote sensing products on a daily

time-step due to the daily fluctuations.

Figure 7.19 AMSR2, SMOS and CRP soil moisture estimates against time

If the daily remote sensing and CRP datasets (Figure 7.19) are converted to a three-day

average (Figure 7.20) this allows the smallest averaging interval, without any gaps in the

dataset. The dataset becomes smoother, as the fluctuations are averaged. This is useful when

comparing the two remote sensing datasets. If the data were averaged over a longer time

interval, the graphs would become smoother.

From Figure 7.20, it can be seen that the remote sensing soil moisture products followed the

general trend of the CRP soil moisture estimates. However, there are still discrepancies in the

validation of the remote sensing products with CRP data, which are due to the horizontal and

vertical scaling.

0

10

20

30

40

50

60

70

3/1/2014 4/30/2014 6/29/2014 8/28/2014 10/27/2014 12/26/2014 2/24/2015

VW

C (

%)

Date

SMOS CRP AMSR2

112

Figure 7.20 Three-day averaged soil moisture estimates

The vertical scaling errors occur, due to the CRP measuring soil moisture at a different depth,

as compared to the remote sensing sensors. The CRP can theoretically measure soil moisture

between depths of 0.12 to 0.72 m, depending on the soil moisture status. The remote sensing

products are a measure of soil moisture at shallow depths (0.00 – 0.05 m).

The horizontal scaling issues occur, as the remote sensing products are at a very coarse

resolution, which greatly exceed the measurement area of the CRP. This is a common

limitation in the validation of remote sensing soil moisture products. However, using the CRP

instead of traditional in-situ techniques, results in an area-averaged estimate instead of a

series of point measurements, which is a great improvement. It is impractical and near

impossible to validate current remote sensing soil moisture products with a traditional in-situ

dataset at the same spatial resolution, due to the heterogeneity of soil moisture.

This limitation cannot currently be overcome in an applied manner, therefore remote sensing

validation has been, and is currently, carried out by validating a “big” area (coarse pixel

value), with a “smaller” area. This will make the coarse pixel area that is not covered by the

validation dataset footprint to be neglected. Thus, true validation cannot be achieved unless

both the remote sensing and validation data are on the same-scale. The term same-scale refers

to an area-averaged soil moisture value and not an in-situ soil moisture network, which

obtains an area-average value by averaging the various in-situ sensors in the network.

0

10

20

30

40

50

60

70

3/1/2014 4/30/2014 6/29/2014 8/28/2014 10/27/2014 12/26/2014 2/24/2015

VW

C (

%)

Date

SMOS CRP AMSR2

113

The SMOS soil moisture dataset seems to have provided better estimates of soil moisture

than the AMSR2 dataset, when compared to the CRP. The SMOS Level Three product has a

coarser resolution (25 km) than the AMSR2 Level Three product (10 km). However, it

provided better estimates of soil moisture. This could be due to the SMOS satellite instrument

using the more recent technology of the L-band wavelength to observe and measure. The

SMOS satellite was launched in 2009 and although the AMSR2 sensor on-board the GCOM-

W1 satellite was launched in 2012, the AMSR2 sensor implements similar instruments and

methods to the previous AMSR on-board the AQUA satellite, for the purpose of data

continuation. Soil moisture retrievals using lower microwave frequencies are expected to be

more accurate (Kim et al., 2015). Thus, although the SMOS satellite was launched before the

GCOM-W1, the SMOS satellite operates at a 1.4 GHz, whilst the AMSR2 sensor operates on

10.7 GHz. The underestimation of soil moisture estimates by the AMSR2 soil moisture

product was also observed in a study by Kim et al. (2015).

The AMSR2 dataset is a more reliable dataset in terms of providing a more complete dataset,

as there are fewer missing days. The daily fluctuations in soil moisture matched the CRP

more than the SMOS, and this is seen by the large variance in the SMOS dataset. In the dry

periods, the AMSR2 dataset did not fluctuate nearly as much as the SMOS dataset.

7.5 SAHG Soil Moisture Product Validation

The SAHG soil moisture product is on a 12.5 x 12.5 km (roughly) spatial grid, which results

in a pixel area of 156.25 km2. In order to obtain a year-long dataset, 2920 images were

downloaded and used to create 365 daily images. The SAHG dataset is continuous and

possesses no gaps in the dataset. The SAHG soil moisture was obtained in SSI and converted

to VWC by using a representative porosity value, which was calculated from the bulk density

values. A time series analysis graph of the SAHG and CRP soil moisture estimates were

plotted (Figure 7.21).

From Figure 7.21, it can be seen that the SAHG product followed the same seasonal trend as

the CRP. There seems to be a close correlation between the datasets, in terms of general

increases and decreases in soil moisture content. The CRP has more variation in soil moisture

from day-to-day. The SAHG product has gradual changes in soil moisture and does not

exhibit the same degree of temporal fluctuation, as seen by the CRP. This difference between

114

the datasets is easily observed. The SAHG product does fluctuate, but the occurrence of these

fluctuations is less than that presented by the CRP dataset. The first nine months (march 2014

to November 2014) of the twelve month time series analysis shows a close correlation

between the CRP and SAHG soil moisture estimates. The last three months (December 2014

to February 2015) shows less of a correlation, as the fluctuations of the CRP soil moisture

estimates do not correspond to the fluctuations of the SAHG soil moisture estimates. This

could be due to an error in the PyTOPKAPI model, such as an error in the input data.

Figure 7.21 Time series analysis of SAHG and CRP soil moisture estimation

A day in summer (7 March 2014) and a day in winter (18 August 2014) are illustrated in

Figures 7.22 and 7.23 respectively. The day in summer showed that the eastern side of South

Africa generally had a higher soil moisture content than the western side, due to eastern side

experiencing most of its rainfall in summer and having a higher mean annual precipitation.

The soil moisture below is expressed as a SSI (%). The study area is adjacent to Lesotho,

which is not covered by this product and therefore has no soil moisture values. In the winter

period, the western side of South Africa increased in soil moisture, whilst the eastern side

decreased, due to the seasonal rainfall patterns. There are still areas in the eastern side of

South Africa that have a high soil moisture content.

0

10

20

30

40

50

60

70

0

10

20

30

40

50

60

70

80

90

100

3/1/2014 5/1/2014 7/1/2014 9/1/2014 11/1/2014 1/1/2015

VW

C (

%)

Rai

nfa

ll (m

m)

Date

Rainfall CRP SAHG

115

Figure 7.22 SAHG daily soil moisture (summer)

116

Figure 7.23 SAHG daily soil moisture (winter)

117

A scatter graph of the CRP (x-axis) was plotted against the SAHG product (y-axis) (Figure

7.24). From Figure 7.24, it can be seen that there is a positive and good correlation between

the datasets. The points are clustered around the 1:1 line (red) and there are no extreme

outliers. There are points below and above the 1:1 line, with the majority of the points being

just below the 1:1 line. The R2 is 0.624, which indicates a close linear relationship between

the datasets. The slope is 1.049, which is close to 1. The points, which are noticeably above

and below the 1:1 line can be attributed to the last three months of the datasets, where there

was less of a correlation between the CRP and SAHG soil moisture estimates. Overall, the

close correlation corresponds to the visual inspection of the data.

Figure 7.24 Scatter graph of CRP against SAHG soil moisture estimates

A graph of the residuals was then plotted against time (Figure 7.25). The ∆ VWC (%) is the

difference between the CRP and SAHG datasets. From Figure 7.25, it is evident that the

majority of the residuals are positive, which indicates that the SAHG soil moisture product

underestimates soil moisture for the majority of the study period. In the dry periods, the

residuals are positive, whilst in the wet periods, there are both positive and negative residuals.

The residuals are generally lower in absolute value in the dry periods, compared to the wet

periods. Once more, the decline in the correlation between the datasets is noticeable between

the first nine months of the study period and the last three months of the study period, which

has the highest residuals with regards to the absolute value of the residuals.

y = 1.0487x - 4.2147R² = 0.6242

0

10

20

30

40

50

60

70

0 10 20 30 40 50 60 70

SAH

G V

WC

(%

)

CRP VWC (%)

118

Figure 7.25 A residual graph of CRP against SAHG was plotted against time

A T-test was conducted. The CRP was variable one and the SAHG product was variable two

(see Table 7.4). From Table 7.4, it can be seen that the means of both variables are similar,

with the CRP mean (33.333) being slightly higher than the SAHG mean (30.742). The

variance of the CRP is 50.292, which is lower than that of the SAHG product (88.607). There

are 677 degrees of freedom. The t stat value is 4.201, which is larger than the t critical two-

tail value of 1.964 and the p value is 3x10-05, which is larger than the alpha value (0.05).

Therefore, according to the T-test, there is a significant difference between the data sets.

Table 7.4 T-test of CRP against SAHG estimates

CRP SAHG Mean 33.333 30.742 Variance 50.292 88.607 Observations 365 365 Hypothesized Mean Difference 0 df 677 t Stat 4.201 P(T<=t) two-tail 3x10-05 t Critical two-tail 1.964

The SAHG soil moisture product provided good estimates of soil moisture, which correlated

well with the CRP estimates. The SAHG soil moisture product measures the SSI (%) in the A

and B soil horizons. In this case, the SSI was obtained at an average depth of one meter.

Therefore, there is a vertical scaling issue, due to the CRP measuring at a depth of around

-25

-20

-15

-10

-5

0

5

10

15

20

25

3/1/2014 5/1/2014 7/1/2014 9/1/2014 11/1/2014 1/1/2015

∆ V

WC

(%)

Date

119

0.12 m, which falls in the range of the SAHG estimate. However, the SAHG product

measures above (0.0 to 0.12 m) and below (0.12 to 1.00 m) and provides an average estimate

of soil moisture. There is a horizontal scaling issue, as the SAHG product is on a 12.5 x 12.5

km spatial grid, which greatly exceeds the measurement footprint of the CRP. For this

research project, all the soil moisture products were converted to volumetric water content

(%), if they were not in volumetric water content (%) already. The SAHG product was in a

SSI and the conversion required an accurate representative porosity value of the study area.

The value determined may not have been the most representative of the study area, but rather

an average of the points selected. This value would result in a constant error throughout the

SAHG volumetric water content dataset.

7.6 SEBS Soil Moisture Validation

In total, 16 relative evaporation and 16 evaporative fraction maps were generated, using the

SEBS Model in ILWIS 3.8.3. These maps were exported, opened and analyzed in ArcGIS

9.3, where the relative evaporation and evaporative fraction of the area within Catchment VI

was determined. The relative evaporation and evaporative fraction values, which were

obtained from the SEBS Model are shown in Table 7.5.

From Table 7.5, it can be seen that the relative evaporation and evaporative fraction results

follow a seasonal trend, as the values are high in summer (wet period) and very low in winter

(dry period), with the intermediate values between the wet and dry periods. Of the 16

estimated relative evaporation and evaporative fraction values, six have high relative

evaporation values (above 0.7), two have intermediate relative evaporation values and eight

have very low relative evaporation values (close to 0).

Catchment VI was burnt on the 5th of September 2014, which could be the reason for the zero

values of the relative evaporation and evaporative fraction being estimated by the SEBS

Model on the 6th of September 2014, 22nd of September 2014 and the 8th of October 2014.

The fire would have resulted in an increase in albedo and a decrease in NDVI, which could

have resulted in these low values. NDVI and albedo are amongst the parameters that the

SEBS Model is most sensitive to (Gibson et al., 2013). A range (high, intermediate and low)

of relative evaporation and evaporative fraction are illustrated in Figure 7.26 and Figure 2.27

respectively.

120

Table 7.5 Relative evaporation and evaporative fraction calculated using the SEBS

model for Catchment VI

Date Relative Evaporation Evaporative Fraction 14-Mar-14 0.73396 0.91684 30-Mar-14 0.94717 0.76871 15-Apr-14 0.87145 0.78335 1-May-14 0.29912 0.50633 17-May-14 Cloud Cloud 2-Jun-14 0.00000 0.00000 18-Jun-14 0.00143 0.00367 4-Jul-14 0.02200 0.04683 20-Jul-14 0.00902 0.01190 5-Aug-14 0.00067 0.00016 21-Aug-14 Cloud Cloud 6-Sep-14 0.00000 0.00000 22-Sep-14 0.00000 0.00000 8-Oct-14 0.00000 0.00000 24-Oct-14 0.11804 0.17727 9-Nov-14 0.93836 0.83581 25-Nov-14 Cloud Cloud 11-Dec-14 0.85773 0.87569 27-Dec-14 Cloud Cloud 12-Jan-15 0.930481 0.94974 28-Jan-15 Cloud Cloud 13-Feb-15 Cloud Cloud

From Figure 7.26, it can be seen that the estimated relative evaporation varies over time. The

30th of March 2014 represents a day when the relative evaporation was very high (0.9471),

the 01st of May represents a day when the relative evaporation was intermediate 0.2991 and

the 05th of August represents a day when the relative evaporation was very low 0.0067.

121

Figure 7.26 A range of different relative evaporation images for Catchment VI

From Figure 7.27, it can be seen that the estimated evaporative fraction varies over time. The

30th of March 2014 represents a day when the evaporative fraction was high (0.7687), the 01st

of May represents a day when the evaporative fraction was intermediate 0.5063 and the 05th

of August represents a day when the evaporative fraction was very low 0.0002.

Catchment VI

122

Figure 7.27 A range of different evaporative fraction images

Total evaporation consists of water that is evaporated from the land surface and water that is

transpired from the vegetation. In order to estimate the actual soil moisture, the saturated soil

moisture content is required, which can be inferred from the porosity. The porosity can be

estimated from the bulk density.

The calculation of soil moisture from relative evaporation and evaporative fraction, estimates

the soil moisture in the root zone, from which the evaporated water (soil evaporation,

transpiration and interception) is sourced. The rooting zone of the grassland vegetation is 0.5

Catchment VI

123

m. Therefore, the average bulk density of the soil from 0.0 to 0.5 m is required to obtain the

porosity of the rooting zone. The bulk density was estimated to be 0.688 g/cm3, therefore the

porosity was calculated to be 0.74. This porosity would be used as the saturated soil moisture

content in the rooting zone.

Two methods for soil moisture estimation were investigated in this component of the research

study. The first was developed by Su et al. (2003) and requires the relative evaporation

(Equation 4.12). The second equation was developed by Scott et al. (2003) and requires the

evaporative fraction (Equation 4.13). The estimated soil moisture was plotted against the

corresponding CRP estimates, which were changed to match the 16-day time-step (Figure

7.28).

From Figure 7.28, it can be seen that both methods follow the same trend, as both methods

generally over-estimate soil moisture in the wet period and generally under-estimate soil

moisture in the dry period. Both methods follow a seasonal trend. The Scott et al. (2003)

method seems to perform better, as it over-estimates less in the wet periods, when compared

to the CRP and under-estimates less in the dry period, compared to the CRP, in comparison to

the Su et al. (2003) method.

Figure 7.28 Time series of CRP estimates and soil moisture back-calculated from the

SEBS model

0

10

20

30

40

50

60

70

80

3/1/2014 4/30/2014 6/29/2014 8/28/2014 10/27/2014 12/26/2014 2/24/2015

VW

C (

%)

Date

CRP (daily) CRP (16-day) Su et al., 2003 Scott et al., 2003

124

Scatter graphs of the CRP against Su et al. (2003) and CRP against Scott et al. (2003) were

then plotted (Figure 7.29). From the scatter graphs in Figure 7.29, it can be seen that both

have positive correlations with points above and below the 1:1 line. The points do not cluster

around the 1:1 line. The points above and below the 1:1 indicate over-estimation and under-

estimation of the methods, when compared with the CRP. Both methods have a similar R2

value but different slopes. The (Scott et al., 2003) method has a R2 value of 0.724 and a slope

of 2.819. The (Su et al., 2003) method has a R2 of 0.722 and a slope of 3.862.

Figure 7.29 Scatter graphs of CRP against the SEBS Model estimates a) CRP against

Su et al. (2003) and b) CRP against Scott et al. (2003).

A graph of the residuals for both the CRP against the Scott et al. (2003) method and the CRP

against the Su et al. (2003) method are illustrated in Figure 7.30. From Figure 7.30, both

residuals follow the same trend. The Su et al. (2003) method has greater residuals as it over-

estimates and under-estimates the most in the respective periods, when compared to the CRP

measurements. The Scott et al. (2003) method, resulted in lower absolute valued residuals

than that of Su et al. (2003).

y = 2.8195x - 65.66R² = 0.7238

0

20

40

60

80

0 20 40 60 80

Sco

tt e

t al

. (2

00

3)

VW

C (

%)

CRP VWC (%)

(a)

y = 3.8621x - 98.376R² = 0.7223

0

20

40

60

80

0 20 40 60 80

Su e

t al

. (2

00

3)

VW

C (

%)

CRP VWC (%)

(b)

125

Figure 7.30 Residual graph of CRP against Su et al. (2003) and Scott et al. (2003).

The results show that the method of back-calculating soil moisture, using both equations,

resulted in an over-estimation of soil moisture in the wet periods and an under-estimation of

soil moisture in the dry period. This could be due to either the estimation of the relative

evaporative and the evaporative fraction in the SEBS Model, or the equations used to

calculate soil moisture from the relative evaporative and the evaporative fraction.

The evaporative fraction estimated using the SEBS Model can be validated with an

evaporative fraction calculated from the eddy covariance system, which is situated in

Catchment VI. The relative evaporation cannot be calculated using the eddy covariance

system, as the relative evaporation requires the determination of Hwet (refer to Equations 4.4

and 4.6), which can only be reliably determined through field experiments to determine wet

and dry bulb temperatures. Thus, the following will only look at determining soil moisture

using the Scott et al. (2003) method, as only the evaporative fraction can be determined from

the eddy covariance system.

The eddy covariance with infrared gas analyser measures latent and sensible heat directly.

The eddy covariance system was operational from the 12th of July 2014 to present. The

evaporative fraction was calculated for the same days as the SEBS estimates, using Equation

8.4 (Bastiaanssen et al., 1997):

-40

-30

-20

-10

0

10

20

30

40

3/14/2014 5/14/2014 7/14/2014 9/14/2014 11/14/2014

∆ V

WC

(%)

Date

Su et al. (2003) Scott et al. (2003)

126

ᴧ = 𝐿𝐸

𝐿𝐸+𝐻 (8.4)

where LE is the latent-heat flux and H is the sensible-heat flux. The eddy covariance system

is on a half-hourly time-step. Therefore, a daily relative evaporation value had to be

determined. The evaporative fraction values determined from the SEBS Model and the eddy

covariance system are seen in Table 7.6 below.

Table 7.6 Evaporative fraction estimates from the SEBS Model and eddy covariance

technique

Date SEBS ᴧ EC ᴧ

14-Mar-14 0.91684 -

30-Mar-14 0.76871 -

15-Apr-14 0.78335 -

1-May-14 0.50633 -

2-Jun-14 0.00000 -

18-Jun-14 0.00367 -

4-Jul-14 0.04683 -

20-Jul-14 0.01190 0.17041

5-Aug-14 0.00016 0.21135

6-Sep-14 0.00000 0.14494

22-Sep-14 0.00000 0.26370

8-Oct-14 0.00000 0.46655

24-Oct-14 0.17727 0.62393

9-Nov-14 0.83581 0.73935

11-Dec-14 0.87569 0.70615

12-Jan-15 0.94974 0.75000

The SEBS Model is under-estimating the evaporative fraction in the dry periods and over-

estimating evaporative fraction in the wet periods, as shown in Table 7.6. It is likely that the

SEBS model was the cause of the poor soil moisture results, rather than the two equations

used. The soil moisture was then estimated by using the calculated eddy covariance derived

evaporative fraction in the (Scott et al., 2003) method and plotted with the CRP

measurements against time (see Figure 7.31).

From Figure 7.31, it can be seen that using the evaporative fraction data from the eddy

covariance system, yielded better results of soil moisture, when compared to the CRP

measurements. Using the eddy covariance data in the Scott et al. (2003) method resulted in

an under-estimation of soil moisture in the dry period, but in the wet period, promising soil

moisture estimates were obtained.

127

Figure 7.31 Time series of the CRP soil moisture estimates and soil moisture from the

Scott et al. (2003) method using the evaporative fraction from SEBS and

the eddy covariance method.

When the evaporative fraction of the SEBS Model is compared to the eddy covariance

estimates of evaporative fraction, it is seen that the SEBS model overestimates evaporative

fraction in the wet period, but under-estimates in the dry period. This indicates the limitations

of the SEBS model. The SEBS model was originally created for agriculture, therefore some

model parameterization is not appropriate for non-agricultural land-covers (Gibson et al.,

2013).

The method proposed by Scott et al. (2003) provided better soil moisture results than the

method proposed by Su et al. (2003). This is mainly due to the Su et al. (2003) equation

linking relative soil moisture to relative evaporation by a linear relationship. The use of the

eddy covariance evaporative fraction estimates in the Scott et al. (2003) method, yielded

good results in the wet period. This technique seems promising due to its spatial resolution.

The technique is not limited temporally, but the choice of satellite (Landsat 8) data,

introduces a temporal limitation. This temporal resolution made this method impractical as a

means to monitor soil moisture, as soil moisture is highly variable over time.

0

10

20

30

40

50

60

70

3/1/2014 4/30/2014 6/29/2014 8/28/2014 10/27/2014 12/26/2014

VW

C (

%)

Date

CRP (16-day) EC-Scott et al. (2003) SEBS-Scott et al. (2003)

128

7.7 Evaluating the Satellite-Based Soil Moisture Products

The satellite-based soil moisture products used in this study all have different spatial

resolutions, as seen in Table 7.7. The SEBS back-calculated estimate, obtains soil moisture

on a 0.09 km2 spatial resolution. This resolution is exceptionally small, compared to the other

products and is the only product that has a spatial resolution smaller than the catchment area.

The AMSR2 soil moisture product has the next best spatial resolution, followed by the

SAHG product. The SMOS product has the largest spatial resolution.

Table 7.7 Spatial resolution of soil moisture products

Product Grid (km) Spatial Resolution (km2) AMSR2 10 x 10 100 SMOS 25 x 25 625 SAHG 12.5 x 12.5 156 SEBS 0.3 x 0.3 0.09

The spatial resolution is a key factor in remote sensing, as there is a need to obtain parameter

data on the finest resolution possible. The temporal resolution is just as important, as the need

to continuously monitor parameters is essential. Thus, smaller temporal resolutions are

required. If the observational days of the satellite-based soil moisture products used in this

study are considered, all the products, except for the SAHG product, have days with no data.

The various soil moisture datasets were then compared on their temporal characteristics, over

the one year study period (Figure 7.32).

From Figure 8.32, it can be seen that the SAHG product has a continuous dataset. The

AMSR2 product is a “daily” product, but consists of a few missing days each month, as the

product does not cover the entire surface of the earth each day. The SMOS data, which had

the largest spatial resolution, also had a lot of missing days each month. This was due to the

product not covering the entire earth’s surface each day, as well as missing pixel values

within the covering bands. The SEBS back-calculated product had the fewest observational

days, as it uses Landsat 8 data, which works on a 16-day interval. This results in a maximum

of two images per month. There was a problem with the sensor, which resulted in no

observations in the months of January and February.

129

Figure 7.32 Number of observation days per month that data was available for each

product

* * * * *

This section detailed the results of the various validations undertaken in the study. The

calibrated CRP soil moisture data set was validated with the representative in-situ soil

moisture data set. The results indicated a close correlation between the data sets and thus

indicated that the CRP is suitable in providing spatial estimates of soil moisture in Cathedral

Peak Catchment VI. The validated CRP soil moisture data set was then used to validate

remote sensing and modelled soil moisture products. The validation of the AMSR2 and

SMOS soil moisture products with the CRP estimates indicated that the AMSR2 and SMOS

products generally underestimated soil moisture but followed the seasonal trend in soil

moisture fluctuation. The SAHG soil moisture product showed the closest correlation when

validated against the CRP estimates. The back-calculation of soil moisture using the relative

evaporation and evaporative fraction, from the SEBS model did not correlate well with the

CRP estimates. There was a general underestimation in winter and overestimation in summer

when the back-calculated soil moisture estimates were compared to the CRP estimates. The

use of the evaporative fraction from the eddy covariance in the method proposed by Scott et

al. (2003), resulted in less of an underestimation in winter and a close correlation in summer

when compared to the CRP estimates. The conclusions drawn from the results and discussion

are detailed in the ensuing chapter.

0

5

10

15

20

25

30

Day

sDAYS SAHG AMSR2 SMOS SEBS

130

8. CONCLUSIONS AND RECOMMENDATIONS

The conclusions and recommendations of the research study are detailed in this chapter.

8.1 Conclusions

Understanding the spatial and temporal variability of soil moisture at different scales is of

great importance in many land surface disciplines, such as hydrology. Soil moisture is a key

hydrological variable, as it impacts the water and energy balance at the land surface-

atmosphere interface and is the main water source for natural vegetation and agriculture. It is

difficult to quantify and assess the soil moisture content at an intermediate scale, due to the

heterogeneity in soil and land cover properties, climate drivers and topography.

The objective of this research study was to compare CRP and satellite-based soil moisture

estimates within Catchment VI of the Cathedral Peak Research Catchments. The three current

methods used to estimate soil moisture were evaluated in this study. These three methods are

in-situ, remote sensing and modelling. Although each method has its own advantages in

measuring soil moisture, they are also greatly limited, as they do not provide soil moisture at

an intermediate scale, which is required for hydrological applications.

The CRP is a new and innovative in-situ instrument that is capable of measuring soil

moisture at an intermediate scale. The CRP, once properly calibrated, is suitable for

providing spatial estimates of soil moisture, as the measurements correlated well with the

representative in-situ soil moisture dataset. The CRP calibration procedure is adequate,

however potential errors can be introduced throughout the procedure, which range from

selecting the sample points, to determining a representative bulk density, to determining the

average neutron count (No) value. Therefore, proper procedure must be adhered to, in order to

minimize potential errors.

The validation of the CRP with a representative catchment soil moisture dataset, from the in-

situ soil moisture network, showed that the CRP is suitable in providing continuous spatial

soil moisture estimates. The in-situ soil moisture dataset could have been more representative

131

of the area, if the echo probes were not destroyed in the fire, however, the remaining in-situ

sensors did create a valuable dataset for the CRP validation.

The calibrated CRP measurements were then used to validate the AMSR2 and SMOS Level

Three soil moisture products. The AMSR2 dataset followed the trend of the CRP and

correlated fairly well, but underestimated soil moisture throughout the study period. The

dataset did have a few missing days. The AMSR2 product was analysed and it was found that

the ascending and descending values correlated well in the dry periods, whilst greater

differences were observed between ascending and descending values in the wet periods.

The SMOS dataset generally under-estimated soil moisture. The dataset did correlate well

and follow the same trend of the CRP. The SMOS dataset had a larger variance and therefore

its values fluctuated more. The dataset had many missing days of data. The SMOS product

was analysed and it was seen that the ascending and descending values were different

throughout. The key issue in the validation of the AMSR2 and SMOS soil moisture products

with the CRP measurements, is the vertical and horizontal scaling issue. Although the CRP is

an improvement from validating remote sensing products with in-situ point measurements,

the difference in measurement depths and the footprint of the CRP and current remote soil

moisture products still remain a limitation.

The AMSR2 and SMOS products consist of ascending and descending values. It has been

shown that there were differences in these values, which could be attributed to changes in soil

moisture in the 12-hour interval between ascending and descending acquisition. It was noted

that there were also differences in day and night geo-physical conditions, with nocturnal

conditions being more favourable for soil moisture retrieval from AMSR2 and SMOS

products.

The CRP was then used to validate modeled soil moisture estimates. These included the

SAHG soil moisture product and the back-calculation of soil moisture from relative

evaporation estimated from the SEBS Model.

The SAHG soil moisture product was validated with the CRP. There was a close correlation

between the SAHG and CRP datasets. The SAHG soil moisture followed the same seasonal

trend as the CRP and had a continuous dataset (no missing values). The SAHG product

generally had a slight underestimation of soil moisture. Although the SAHG product

132

performed well, there was still the presence of vertical and horizontal scaling issues, due to

differences in the measurement depth and footprint of the two datasets. There is also the issue

of the conversion of the SSI to VWC, which required a representative porosity of the study

area to be determined.

The back-calculation of soil moisture from relative evaporation and evaporative fraction,

estimated using the SEBS model, looked like a promising technique. The spatial resolution

was less than the catchment area and the measurement depth was representative of the root

zone of the vegetation (0.50 m). Therefore, this product would have the least horizontal and

vertical scaling issues, when validated against the CRP. The SEBS model did not provide

accurate estimates of relative evaporation and evaporative fraction compared to the

evaporative fraction estimates from the eddy covariance system. The SEBS Model over-

estimated the evaporative fraction in the wet period, whilst under-estimating the evaporative

fraction in the dry periods, which resulted in soil moisture over-estimation in the wet periods

and under-estimation in the dry periods. When the evaporative fraction derived from the eddy

covariance system was used in the (Scott et al., 2003) method, the resultant soil moisture

estimates, under-estimated in the dry period, however promising estimates were obtained in

the wet period, which correlated well with the CRP estimates. Although the back-calculation

method results in soil moisture estimates on a 30 m spatial grid, the temporal resolution of the

imagery used is 16 days, which is very impractical for continuous soil moisture monitoring.

8.2 Recommendations

The following recommendations can be used to address the main limitations that were

experienced in this research study. This will provide assistance for future research studies.

i. The calibration of the CRP is both time and labour intensive, as calibrations over

different periods are required. There are a variety of errors that could emerge

throughout the calibration procedure. Therefore, the proper techniques and

equipment must be used to minimize the occurrence of errors, in order to obtain

reliable soil moisture measurements. The use of a TDR HydroSense probe, which

obtains instantaneous measurements of soil moisture, when inserted into the soil, can

potentially be used to obtain the necessary soil moisture measurements for the

calibration. Thus, it would greatly reduce the time and labour required.

133

ii. The validation of the CRP needs to be done with in-situ soil moisture sensors at a

variety of depths. Ideally several TDR pits, in the CRP measurement volume could

be used. This would improve the validation, as the CRP doesn’t measure at a constant

depth and therefore, shouldn’t be validated with in-situ measurements at a constant

depth. However, this would be very capital, time and labour intensive.

iii. The potential for using temporal stability analysis to find a more representative

sensor location would be useful to upscale the in-situ observations to the entire

catchment.

iv. Current remote sensing soil moisture products are still too coarse to be validated with

CRP measurements. Although the use of the CRP for the validation of remote

sensing soil moisture products is an improvement, the vertical and horizontal scaling

issues still remain as major limitations. This limitation could potentially be addressed

through the downscaling of remote sensing soil moisture products, in order to obtain

finer spatial scale soil moisture estimates.

v. The back-calculation of soil moisture using relative evaporation and evaporated

fraction, from the SEBS Model, indicated that the use of the SEBS Model is not

suitable in the study area for the estimation of these parameters. Future research

should limit the use the SEBS model to agricultural landscapes.

134

9. REFERENCES

Albergel, C, de Rosnay, P, Gruhier, C, Mu˜noz-Sabater, J, Hasenauer, S, Isaksen, L, Kerr, Y

and Wagner, W. 2012. Evaluation of remotely sensed and modelled soil moisture

products using global ground-based in situ observations. Remote Sens. Environ 118:

215-226.

Allen, R, Bastiaanssen, W and Waters, R. 2002. Personal communication, the University of

Idaho and the Idaho Department of Water Resources, The University of Idaho,

Bastiaanssen, W, Pelgrum, H, Droogers, P, de Bruin, H and Menenti, M. 1997. Area-average

estimates of evaporation, wetness indicators and top soil moisture during two golden

days in EFEDA. Agric Forest Meteorol 87: 119-137.

Bezerra, B, dos Santos, C, da Silva, B, Bezerra, M, Bezerra, J and Rao, T. 2013. Estimation

of soil moisture in the root-zone from remote sensing data. . Ci. Solo 37: 596-603.

Brocca, L, Hasenauer, S, Lacava, T, Melone, F, Moramarco, T, Wagner, W, Dorigo, W and

Bittelli, M. 2011. Soil moisture estimation through ASCAT and AMSR-E sensors: An

intercomparison and validation study across Europe. Remote Sensing of Environment

115: 3390-3408.

Brocca, L, Tarpanelli, A, Moramarco, T, Melone, F, Ratto, S, Cauduro, M, Ferraris, S, Berni,

N, Ponziani, F, Wagner, W and Melzer, T. 2013. Soil moisture estimation in Alpine

Catchments through modeling and satellite observations. Soil Science Society of

America

Desilets, D and Zreda, M. 2013. Footprint diameter for a cosmic-ray soil moisture probe:

theory and Monte Carlo simulations. Water Resources Research 49: 3566-3575.

Desilets, D, Zreda, M and Ferre, T. 2010. Nature’s neutron probe: land surface hydrology at

an elusive scale with cosmic rays. Water Resources Research 46: 1-7.

Dorigo, W, Wagner, W, Hohensinn, H, Paulik, C, Xaver, A and Jackson, T. 2011. The

International Soil Moisture Network: A data hosting facility for global in situ soil

moisture measurements. Hydrology and Earth System Sciences. Hydrol. Earth Syst.

Sci 15: 1675-1698.

Draper, CS, Walker, JP, Steinhle, P and Holmes, TRH. 2009. An evaluation of AMSR–E

derived soil moisture over Australia. Remote Sensing of Environment 113 (4): 703-

710.

135

Dutta, R and D'este, C. 2013. Virtual Calibration of Cosmic Ray Sensor: using supervised

ensemble machine Learning. IJACSA) International Journal of Advanced Computer

Science and Applications 4: (8): 104-110.

Engman, ET. 1991. Application of microwave remote sensing of soil moisture for water

resources and agriculture. Remote Sensing of Environment 35: 213-226.

Everson, CS, Molefe, GL and Everson, TM. 1998. Monitoring and modelling components of

the water balance in a grassland catchment in the summer rainfall area of South

Africa.

Fang, B and Lakshmi, V. 2014. Soil moisture at watershed scale : Remote sensing techniques

Journal of Hydrology 516: 258-272.

Franz, T. 2014. Personal communication, Franz Hydrogeophysics Lab Group, 1-4.,

Franz, T, Wang, T, Avery, W, Finkenbiner, C and Brocca, L. 2015. Spatiotemporal

characterization of soil moisture fields in agricultural areas using cosmic-ray neutron

probes and data fusion. EGU General Assembly 17

Franz, T, Zreda, M, Ferre, T and Rosolem, R. 2013. An assessment of the effect of horizontal

soil moisture heterogeneity on the area-average measurement of cosmic-ray neutrons.

Water Resources Research 49: 1-9.

Franz, T, Zreda, M, Rosolem, R and Ferre, T. 2012a. Field Validation of a cosmic-ray

neutron sensor using a distributed sensor network. Vadose Zone Journal 11

Franz, T, Zreda, M, Rosolem, R and Ferre, T. 2012b. A universal calibration function for

determination of soil moisture with cosmic-ray neutrons. Hydrology and Earth System

Sciences Discussions 9: 10303-10322.

Gibson, L, Jarmain, C, Su, Z and Eckardt, F. 2013. Estimating evapotranspiration using

remote sensing and the Surface Energy Balance System - A South African

Perspective. Water SA 39: 477-484.

Gokman, M, Vekerdy, Z, Verhoef, A, Verhoef, W, Batelaan, O and van der Tol, C. 2012.

Integration of soil moisture in SEBS for improving evapotranspiration estimation

under water stress conditions. Remote Sensing of Environment 121: 261-274.

Gruber, A, Dorigo, WA, Zwieback, S, Xaver, A and Wagner, W. 2013. Characterizing

coarse-scale representativeness of in situ soil moisture measurements from the

International soil moisture network. gsvadzone 12 (2):

Gruhier, C, Albergel, C, de Rosnay, P and Zeiner, B. 2011. Comparison between H-SAF

large scale surface soil moisture, H-SAF assimilated soil moisture and SMOS Level 2

soil moisture. CESBIO, Toulouse, France.

136

Guillem, M. 2010. Multi-scale soil moisture retrievals from microwave remote sensing

observations. Unpublished thesis, Remote Sensing Lab, Dept. Teoria del Senyal I

Comunicacions Universitat Polit Ecnica de Catalunya,

Gush, M, Scott, DF, Jewitt, GPW, Schulze, RE, Lumsden, T, Hallowes, L and Gorgens, A.

2002. Estimation of streamflow reduction resulting from commercial afforestation in

South Africa.

Hawdon, A, Mcjannet, D and Wallace, J. 2014. Calibration and correction procedures for

cosmic-ray neutron soil moisture probes located across Australia. Water Resources

Research 50 (6): 5029-5043.

Hillel, D. 2008. In: ed. Hillel, D, Soil in the Environment. Academic Press, San Diego.

Imaoka, K, Kachi, M, Kasahara, M, Ito, N, Nakagawa, K and Oki, T. 2010. Instrument

performance and calibration of AMSR-E and AMSR2. International Archives of the

Photogrammetry, Remote Sensing and Spatial Infromation Science 38: 13-16.

Jackson, TJ, Cosh, MH, Bindlish, R, Starks, PJ, Bosch, DD, Seyfried, M, Goodrich, DC,

Moran, MS and Du, J. 2010. Validation of advanced microwave scanning radiometer

soil moisture products. IEEE Transactions on Geoscience and Remote Sensing 48

(12):

Jarmain, C, Govender, M and Everson, CS. 2004. Improving the basis for predicting total

evaporation from natural veld types in South Africa.

JAXA, JAEA. 2013a. AMSR Level 3 Product Format Description Document.

JAXA, JAEA. 2013b. GCOM-W1 "SHIZUKU" Data Users Handbook. First Edition.

Jiao, Q, Zhu, Z and Du, F. 2014. Theory and application of measuring mesoscale soil

moisture by cosmic-ray fast neutron probe. 35th International Symposium on Remote

Sensing of Environment 17

Jones, SB, Wraith, JM and Or, D. 2002. Time domain reflectometry measurement principles

and applications. Hydrological Processes 16: 141-153.

Kerr, YH, Waldteufel, P, Richaume, P, Wigneron, JP, Ferrazzoli, P, Mahmoodi, A, Al Bitar,

A, Cabot, F, Gruhier, C, Juglea, SE, Leroux, D, Mialon, A and Delwart, S. 2012. The

SMOS soil moisture retrieval algorithm. Geoscience and Remote Sensing, IEEE

Transactions on 50 (5): 1384-1403.

Kerr, YH, Waldteufel, P, Wigneron, JP, Delwart, S, Cabot, F, Boutin, J, Escorihuela, MJ,

Font, J, Reul, N, Gruhier, C, Juglea, SE, Drinkwater, MR, Neira, M and Mecklenburg,

S. 2010. The SMOS mission: new tool for monitoring key elements of the global

water cycle. Proceedings of the IEEE 98 (5): 666-687.

137

Kim, S, Liu, YY, Johnson, FM, Parinussa, RM and Sharma, A. 2015. A global comparison of

alternate AMSR2 soil moisture products: Why do they differ ? Remote Sensing of

Environment

Kohli, M, Shron, M, Zreda, M, Schmidt, U, Dietrich, P and Zacharias, S. 2015. Footprint

characteristics revised for field-scale soil moisture monitoring with cosmic-ray

neutrons. Water Resources Research 10: 5772-5790.

Koike, T, Nakamura, YK and Fujii, I. 2004. Development of an Advanced Microwave

Scanning Radiometer (AMSR-E) Algorithm of soil moisture and vegetation water

content. Annual Journal of Hydraulic Engineering 48 (2): 217-222.

Kuenene, BT, van Huyssteen, CW and Hensley, M. 2009. Soil water saturation in the

Cathedral Peak VI catchment, KwaZulu-Natal. SAGA Biennial Technical Meeting and

Exhibition 11: 565-570.

Lakshmi, V. 2013. Remote sensing of soil moisture. ISRN Soil Science 2013: 33 pages.

Leroux, D, Kerr, YH, Al Bitar, A, Bindlish, R, Jackson, TJ, Berthelot, N and Portet, G. 2013.

Comparison between SMOS, VUA, ASCAT and ECMWF soil moisture products

over four watersheds in United States. IEEE Transactions on Geoscience and Remote

Sensing 1-20.

Lu, H, Koike, T and Yang, K. 2014. Improvement of AMSR2 soil moisture algorithm with

considering temperature profile effects in dry soil: a case study in Heihe Basin. IEEE

3299-3302.

Ma, W, Hafeez, M, Rabbani, U, Ishikawa, H and Ma, Y. 2012. Retrieved actual ET using

SEBS model from Landsat-5 TM data for irrigation area of Australia. Atmospheric

Environment 59: 408-414.

Markham, B, Storey, J and Morfitt, R. 2015. Landsat-8 sensor characterization and

calibration. Remote Sens 2015 (7): 2279-2282.

Mecklenburg, S, Drusch, M, Kerr, Y, Font, J and Roger, O. 2013. ESA’s soil moisture and

ocean salinity mission: mission performance and operations. European Space Agency,

ESA-ESRIN, Via Galileo Galilei, Frascati, Italy.

Mekonnen, D. 2009. Satellite remote sensing for soil moisture estimation: Gumara

catchment, Ethiopia. Unpublished thesis, International institute for geo-information

science and earth observation, Netherlands.

Meng, L and Quiring, SM. 2008. A Comparison of Soil Moisture Models Using Soil Climate

Analysis Network Observations. American Meteorological Society 9: 641-659.

138

Mengistu, MG, Everson, CS, Moyo, NC and Savage, MJ. 2014. The validation of the

variables (evaporation and soil moisture) in hydrometeorological models.

Merlin, O, Al Bitar, A, Walker, JP and Kerr, Y. 2009. A sequential model for disaggregating

near-surface soil moisture observations using multi-resolution thermal sensors.

Remote Sensing of Environment 113 (10): 2275-2284.

Merlin, O, Al Bitar, A, Walker, JP and Kerr, Y. 2010. An improved algorithm for

disaggregating microwave-derived soil moisture based on red, near-infrared and

thermal-infrared data. Remote Sensing of Environment 114 (10): 2305-2316.

Merlin, O, Rudiger, C, Al Bitar, A, Richaume, P, Walker, JP and Kerr, YH. 2012.

Disaggregation of SMOS soil moisture in southeastern Australia. Geoscience and

Remote Sensing, IEEE Transactions on 50 (5): 1556-1571.

Merlin, O, Walker, JP, Chehbouni, A and Kerr, Y. 2008. Towards deterministic downscaling

of SMOS soil moisture using MODIS derived soil evaporative efficiency. Remote

Sensing of Environment 112 (10): 3935-3946.

Ni-Meister, W. 2005. Soil moisture initialization for climate prediction: characterization of

model and observation errors. Journal of Geophysical Research 110: 11-29.

Ochsner, TE, Cosh, MH, Cuenca, RH, Dorigo, WA, Draper, CS, Hagimoto, Y, Kerr, YH,

Larson, KM, Njoku, EG, Small, EE and Zreda, M. 2013. State of the art in large-scale

soil moisture monitoring. . Soil Science Society of America.

Parrens, M, Zakharova, E, Lafont, S, Calvet, JC, Kerr, YH, Wagner, W and Wigeron, JP.

2012. Comparing soil moisture retrievals from SMOS and ASCAT over France.

Hydrol. Earth Syst. Sci 16: 423-440.

Piles, M, Camps, A, Vall-llossera, M, Corbella, I, Panciera, R, Rudiger, C, Kerr, YH and

Walker, JP. 2011. Downscaling SMOS-derived soil moisture using MODIS

visible/infrared Data. IEEE Transactions on Geoscience and Remote Sensing 49 (9):

Qin, J, Yang, K, Chen, Y, Zhao, L and Han, M. 2013. Spatial up-scaling of in-situ soil

moisture measurements based on MODIS-derived apparent thermal inertia. Remote

Sensing of Environment 138: 1-9.

Sabins, FF. 2007. Remote Sensing Principles and Interpretation. Waveland Press, Illinois.

Scott, C, Bastiaanssen, W and Ahmad, M. 2003. Mapping spatio-temporal distributions of

soil moisture throughout irrigated watersheds using optical and high resolution

imagery. ASCE Irrigat and Drain Eng 129

Scott, DF. 1994. The hydrological effects of fire in South African catchments. . Unpublished

thesis, Department of Agricutural Engineering, University of Natal, South Africa,

139

Scott, DF, Prinsloo, FW, Moses, G, Mehlomakulu, M and Simmers, ADA. 2000. A re-

analysis of the South African catchment afforestation experimental data

Shin, Y and Mohanty, BP. 2013. Development of a deterministic downscaling algorithm for

remote sensing soil moisture footprint using soil and vegetation classifications. Water

Resources Research 49: 6208-6228.

Sinclair, S and Pegram, GGS. 2010. A comparison of ASCAT and modeled soil moisture

over South Africa, using TOPKAPI in land surface mode. Hydrol. Earth Syst. Sci. 14:

613-626.

Sinclair, S and Pegram, GGS. 2012. PyTOPKAPI – an open-source implementation of the

TOPKAPI hydrological model. Pegram and Associates.

Sinclair, S and Pegram, GGS. 2013a. HYLARSMET: A hydrologically consistent land surface

model for soil moisture and evapotranspiration modelling over Southern Africa using

remote sensing and meteorological data.

Sinclair, S and Pegram, GGS. 2013b. A sensitivity assessment of the TOPKAPI model with

an added infiltration module. Journal of Hydrology 479: 100-112.

Song, C, Jia, L and Menenti, M. 2013. Retrieving high-resolution surface soil moisture by

downscaling AMSR-E brightness temperature using MODIS LST and NDVI data.

IEEE Journal of Selected Topics in Applied Earth Observations and Remote Sensing

(2013): 1939-1404.

Stoof, CR, Vervoort, RW, Iwema, J, van de Elsen, E, Ferreira, AJD and Ritsema, CJ. 2012.

Hydrological response of a small catchment burned by experimental fire. Hydrol.

Earth Syst. Sci 16: 267-285.

Su, Z. 2002. The Surface Energy Balance System (SEBS) for estimation of turbulent heat

fluxes. Hydrology and Earth System Sciences 6: 85-99.

Su, Z, Yacob, A, Wen, J, Roerink, G, He, Y, Gao, B, Boogaard, H and van Diepen, C. 2003.

Assessing relative soil moisture with remote sensing data: theory, experimental

validation, and application to drought monitoring over the North China Plain. Physics

and Chemistry of the Earth 28: 89-101.

USGS. 2013. Using the USGS Landsat 8 Product. [Internet]. Available from:

http://landsat.usgs.gov/Landsat8_Using_Product.php. [Accessed: February].

USGS. 2015. Landsat 8 (L8) data users handbook

Villarreyes, CAR, Baroni, G and Oswald, SE. 2013. Calibration approaches of cosmic-ray

neutron sensing for soil moisture measurement in cropped fields. Hydrology and

Earth System Sciences 10: 4237-4274.

140

Vischel, T, Pegram, GGS, Sinclair, S and Parak, M. 2008. Implementation of the TOPKAPI

model in South Africa: initial results from the Liebenbergsvlei catchment. Water SA

34 (3): 1-12.

Wagner, W. 2008. Remote sensing of soil moisture in support to hydrological and

meteorological modelling. METIER Training Course "Remote Sensing of the

Hydrosphere" Finish Environment Institute, Helsinki, Finland.

Wagner, W, Dorigo, W, de Jeu, R and Fernandez, JC. 2012. Fusion of active and passive

microwave observations to create an essential climate variable data record on soil

moisture. Vienna University of Technology, Institute of Photogrammetry and Remote

Sensing, Gusshausstrasse, 1040 Wien, Austria.

Walker, JP, Willgoose, GR and Kalma, JD. 2004. In situ measurement of soil moisture: a

comparison of techniques. Journal of Hydrology 2963: 85-99.

Wang, L and Qu, J. 2009. Satellite of remote sensing application for surface soil moisture

monitoring. Front. Earth Sci. 3 (2): 237-247.

Zazueta, FS and Xin, J. 1994. Soil Moisture Sensors. University of Florida.

Zhang, A, Jia, G, Wang, S, Zhoa, TB and Feng, JM. 2011. Evaluation of AMSR-E derived

soil moisture over Northern China. China. Atmospheric and Oceanic Science. 4 (4):

223-228.

Zhao, W and Li, A. 2013. A downscaling method for improving the spatial resolution of

AMSR-E derived soil moisture product based on MSG-SEVIRI data. Remote Sensing

of Environment 5 (2013): 6790-6811.

Zreda, M, Desilets, D, Ferre, TPA and Scott, RL. 2008. Measuring soil moisture content non-

invasively at intermediate spatial scale using cosmic-ray neutrons. Geophysical

Research 35

Zreda, M, Shuttleworth, WJ, Zeng, X, Zweck, C, Desilets, D, Franz, T and Rosolem, R.

2012. COSMOS: the COsmic-ray Soil Moisture Observing System. Hydrol. Earth

Syst. Sci. 16: 4079-4099.